4211 - Water Research Foundation

Transcription

4211 - Water Research Foundation
Lead(IV) Oxide Formation and
Stability in Drinking Water
Distribution Systems
Report #4211
Subject Area: Infrastructure
Lead(IV) Oxide Formation and
Stability in Drinking Water
Distribution Systems
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
About the Water Research Foundation
The Water Research Foundation is a member-supported, international, 501(c)3 nonprofit
organization that sponsors research that enables water utilities, public health agencies, and
other professionals to provide safe and affordable drinking water to consumers.
The Foundation’s mission is to advance the science of water to improve the quality of life.
To achieve this mission, the Foundation sponsors studies on all aspects of drinking water,
including resources, treatment, and distribution. Nearly 1,000 water utilities, consulting
firms, and manufacturers in North America and abroad contribute subscription payments to
support the Foundation’s work. Additional funding comes from collaborative partnerships
with other national and international organizations and the U.S. federal government,
allowing for resources to be leveraged, expertise to be shared, and broad-based knowledge
to be developed and disseminated.
From its headquarters in Denver, Colorado, the Foundation’s staff directs and supports the
efforts of more than 800 volunteers who serve on the board of trustees and various
committees. These volunteers represent many facets of the water industry, and contribute
their expertise to select and monitor research studies that benefit the entire drinking water
community.
Research results are disseminated through a number of channels, including reports, the
Website, Webcasts, workshops, and periodicals.
The Foundation serves as a cooperative program providing subscribers the opportunity to
pool their resources and build upon each others’ expertise. By applying Foundation research
findings, subscribers can save substantial costs and stay on the leading edge of drinking
water science and technology. Since its inception, the Foundation has supplied the water
community with more than $460 million in applied research value.
More information about the Foundation and how to become a subscriber is available at
www.WaterRF.org.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Lead(IV) Oxide Formation and
Stability in Drinking Water
Distribution Systems
Prepared by:
Yin Wang, Yanjiao Xie, and Daniel E. Giammar
Department of Energy, Environmental and Chemical Engineering
Washington University, 1 Brookings Drive, St. Louis, MO 63130
Sponsored by:
Water Research Foundation
6666 West Quincy Avenue, Denver, CO 80235
Published by:
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
DISCLAIMER
This study was funded by the Water Research Foundation (Foundation). The Foundation
assumes no responsibility for the content of the research study reported in this publication or
for the opinions or statements of fact expressed in the report. The mention of trade names
for commercial products does not represent or imply the approval or endorsement of the
Foundation. This report is presented solely for informational purposes.
Copyright © 2012
by Water Research Foundation
ALL RIGHTS RESERVED.
No part of this publication may be copied, reproduced
or otherwise utilized without permission.
ISBN 978-1-60573-180-3
Printed in the U.S.A.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
CONTENTS
LIST OF TABLES ........................................................................................................................ vii
LIST OF FIGURES ....................................................................................................................... ix
FOREWORD ............................................................................................................................... xiii
ACKNOWLEDGMENTS ............................................................................................................ xv
EXECUTIVE SUMMARY ........................................................................................................ xvii
CHAPTER 1: INTRODUCTION .................................................................................................. 1 Significance of Lead(IV) Oxides to Lead in Drinking Water............................................. 1 Project Objectives, Hypotheses, and Structure of the Report ............................................. 2 Project Objectives ......................................................................................................... 2 Structure of the Report .................................................................................................. 3 Background on Lead(IV) Oxides and Relevance to Drinking Water ................................. 4 Lead(IV) Oxide Occurrence and Stability in Distribution Systems ............................. 5 Dissolution Rates of Lead Corrosion Products ............................................................. 8 Mitigation of Lead Release from Corrosion Products on Lead Pipe ............................ 9 CHAPTER 2: FACTORS CONTROLLING THE FORMATION OF LEAD(IV) OXIDES ..... 11 Overview of Research on Lead(IV) Oxide Formation ..................................................... 11 Materials and Methods ...................................................................................................... 11 Materials ..................................................................................................................... 11 PbO2 Formation Experiments ..................................................................................... 12 Analytical Methods ..................................................................................................... 14 Results and Discussion ..................................................................................................... 14 Overview ..................................................................................................................... 14 Products of Reaction with Massicot ........................................................................... 14 Products of Reaction with Elemental Lead(0) Powder ............................................... 18 Products of Reaction with Dissolved Lead Chloride .................................................. 20 Products of Reaction with Lead(II) Carbonates.......................................................... 21 Summary of Lead(IV) Oxide Formation .................................................................... 25 CHAPTER 3: DISSOLVED LEAD CONCENTRATIONS IN SOLUTIONS CONTAINING
LEAD(IV) OXIDE........................................................................................................................ 27 Overview of Lead(IV) Oxide Solubility ........................................................................... 27 Material and Methods ....................................................................................................... 29 Materials ..................................................................................................................... 29 Synthesis of Scrutinyite .............................................................................................. 29 PbO2 Dissolution Experiments ................................................................................... 30 Analytical Methods ..................................................................................................... 31 Results and Discussion ..................................................................................................... 31 Effect of Total Lead Concentration on the Formation of Scrutinyite ......................... 31 v
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
vi | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Lead Release from Plattnerite ..................................................................................... 32 Lead Release from Scrutinyite .................................................................................... 35 Summary of Lead(IV) Oxide Solubility ..................................................................... 36 CHAPTER 4: DISSOLUTION RATES OF LEAD(IV) OXIDE ................................................ 39 Overview of Research on Lead(IV) Oxide Dissolution Rates .......................................... 39 Materials and Methods ...................................................................................................... 39 Materials ..................................................................................................................... 39 PbO2 Dissolution Experiments ................................................................................... 41 Pb(II) Adsorption Experiments ................................................................................... 45 Analytical Methods ..................................................................................................... 45 Results and Discussion ..................................................................................................... 45 Overview ..................................................................................................................... 45 PbO2 Dissolution at Baseline Conditions – Ultrapure Water (Set A) ......................... 46 Effect of Chemical Reductant and Complexing Ligand on PbO2
Dissolution (Set B)...................................................................................................... 54 Effect of Free Chlorine on PbO2 Dissolution (Set C) ................................................. 62 Effect of Orthophosphate on PbO2 Dissolution (Set D) ............................................. 64 Summary of PbO2 Dissolution Rates .......................................................................... 68 CHAPTER 5: SUMMARY AND CONCLUSIONS ................................................................... 69 Summary of Project .......................................................................................................... 69 Formation of Lead(IV) Oxide ........................................................................................... 69 Solubility of Lead(IV) Oxide ............................................................................................ 69 Dissolution Rate of Lead(IV) Oxide ................................................................................. 70 CHAPTER 6: RECOMMENDATIONS TO UTILITIES ........................................................... 71 Determining Whether or not Lead(IV) Oxides are Present .............................................. 71 Maintaining Low Dissolved Lead Concentrations in Distribution Systems
that Contain Lead(IV) Oxides ........................................................................................... 71 APPENDIX A: EQUILIBRIUM CONSTANTS AND REACTIONS........................................ 73 APPENDIX B: SUPPORTING INFORMATION ON LEAD(IV) OXIDE FORMATION ....... 77 APPENDIX C: SUPPORTING INFORMATION ON LEAD(IV) OXIDE DISSOLUTION
RATES .......................................................................................................................................... 85 REFERENCES ............................................................................................................................. 89 ABBREVIATIONS ...................................................................................................................... 93 ©2012 Water Research Foundation. ALL RIGHTS RESERVED.
LIST OF TABLES
2.1
Experimental conditions studied for the formation of PbO2 ..............................................14
2.2
Summary of the resulting sold phases ...............................................................................15
3.1
Reactions and rate constants of PbO2 dissolution ..............................................................27
3.2
Experimental conditions studied for the solubility of PbO2 ..............................................31
4.1
Experimental conditions studied for the dissolution rate of PbO2 .....................................42
4.2
Conditions studied for PbO2 dissolution ............................................................................43
4.3
Conditions and results of plattnerite dissolution rate experiments ....................................47
4.4
Thermodynamics of PbO2 dissolution ...............................................................................52
4.5
Summary of the dissolution rate of PbO2 by different reaductants ...................................60
A.1
Equilibrium constants for aqueous species ........................................................................73
A.2
Solubility products of select lead solids.............................................................................74
A.3
Chemical potentials for various aqueous species...............................................................75
vii
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
LIST OF FIGURES
ES.1
Integrated approach to investigate dynamics of PbO2 in distribution............................ xviii
1.1
Integrated approach to investigate dynamics of PbO2 in distribution..................................3
1.2
Predominance areas of lead(0)/lead(II)/lead(IV) system for 15 μg/L total lead and 30
mg/L dissolved inorganic carbon .............................................................................6
1.3
Dissolved lead concentration in equilibrium with (a) lead(IV) oxide at sufficiently
oxidizing conditions that all lead is present as lead(IV) and (b) the lead(II)
carbonate hydrocerussite with 30 mg/L DIC and the lead(II) phosphate
hydroxylpyromorphite with 30 mg/L DIC and 3 mg/L orthophosphate..................6
2.1
Potential pathways of PbO2 formation from initial Pb(II) phases .....................................11
2.2
X-ray diffraction patterns of materials in this study ..........................................................12
2.3
Electron micrographs showing the morphology of pure (a) cerussite, (b)
hydrocerussite, (c) massicot and (d) lead(0) powder .............................................13
2.4
X-ray diffraction patterns of solids following reaction of massicot with 20 mg Cl2/L
free chlorine in the absence and presence of 20 mg C/L DIC after 28 days..........16
2.5
Electron micrographs of PbO2 products from reaction of PbO (a) after 1 day at pH 7.5
with no DIC and 20 mg Cl2/L free chlorine, and (b) after 1 day at pH 10 with
20 mg/L DIC and 20 mg Cl2/L free chlorine .........................................................16
2.6
X-ray diffraction patterns of solids following reaction of massicot at pH 7.5 with and
without 20 mg C/L dissolved inorganic carbon with different free chlorine
concentrations (in mg/L as Cl2) after 7 and 28 days ..............................................17
2.7
Electron micrographs of solids from massicot reaction at pH 7.5 with 20 mg Cl2/L
free chlorine, no DIC, after 28 days .......................................................................18
2.8
X-ray diffraction patterns of solids following reaction of lead(0) powder at pH 8.5
with 20 mg C/L dissolved inorganic carbon and 20 mg Cl2/L free chlorine
after different reaction time ...................................................................................19
Electron micrographs of PbO2 products from reaction of lead(0) powder at pH 8.5
with 20 mg C/L DIC and 20 mg Cl2/L free chlorine .............................................19
2.9
2.10
X-ray diffraction patterns of solids following reaction of lead(0) powder at pH 8.5
with 20 mg C/L dissolved inorganic carbon with different free chlorine
concentrations (in mg/L as Cl2) after 28 days ........................................................20
ix
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
x | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
2.11
X-ray diffraction patterns of solids following reaction of lead chloride solution in the
absence of dissolved inorganic carbon with different free chlorine
concentrations (in mg/L as Cl2) after 28 days ........................................................21
2.12
X-ray diffraction patterns of solids following reaction of lead chloride solution in the
absence of dissolved inorganic carbon (a) at pH 7.5 with 20 mg Cl2/L free
chlorine and (b) at pH 10 with 4 mg Cl2/L free chlorine .......................................22
2.13
X-ray diffraction patterns of solids following reaction of hydrocerussite in the absence
of dissolved inorganic carbon with different free chlorine concentrations
(in mg/L as Cl2) after 28 days ................................................................................23
2.14
X-ray diffraction patterns of solids following reaction of cerussite in the absence of
dissolved inorganic carbon with different free chlorine concentrations
(in mg/L as Cl2) after 28 days ................................................................................24
2.15
Electron micrographs of the solid from cerussite reaction at (a) pH 10, 20 mg Cl2/L
free chlorine, and no DIC after 1 day, and (b) pH 7.5, 42 mg Cl2/L free chlorine,
and no DIC after 7 days .........................................................................................25
2.16
Formation products from Pb(II) starting phases ................................................................26
3.1
Predicted dissolved lead concentration in equilibrium with scrutinyite in the presence
of 2.8 * 10-5 M HOCl and Cl- ................................................................................28
3.2
X-ray diffraction patterns of scrutinyite synthesized in this study ....................................29
3.3
Electron micrographs showing the morphology of scrutinyite ..........................................30
3.4
X-ray diffraction patterns of solids following reaction of hydrocerussite at pH 10
without DIC and with a free chlorine:lead ratio of 1.2:1 .......................................32
3.5
Dissolved lead concentrations with time in plattnerite batch dissolution experiments .....33
3.6
Predicted and measured equilibrium solubility of PbO2 the presence of 2 mg Cl2/L free
chlorine and with no DIC.......................................................................................34
3.7
Dissolved lead concentrations with time in scrutinyite batch dissolution experiments.....37
4.1
X-ray diffraction patterns of the PbO2 used in this study ..................................................40
4.2
Electron micrographs showing the morphology of pure plattnerite ..................................40
4.3
Flow-through reactor for measuring dissolution rates .......................................................42
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Figures | xi
4.4
Effluent lead concentrations and pH from CSTRs as a function of the number of
hydraulic residence times (tres = 30 min) in the 50 mg C/L DIC-buffered water
at pH (a) 5.7, (b) 6.7, (c) 7.6, and (d) 8.5 ...............................................................52
4.5
The dissolution rate of plattnerite in the 50 mg C/L DIC-buffered water at pH 5.7 to
8.5...........................................................................................................................53
4.6
Adsorption of 0.96 µM Pb(II) onto 0.1 g/L PbO2(s) in the presence of 50 mg C/L
dissolved inorganic carbon (DIC) as a function of pH ..........................................53
4.7
Effect of pH on the dissolution rate of plattnerite with 50 mg C/L DIC in the absence
and presence of 10 μM iodide ................................................................................54
4.8
Effect of iodide concentration on the dissolution rate of plattnerite with 50 mg C/L
DIC at pH 7.6 and 8.5 ............................................................................................55
4.9
X-ray diffraction patterns of the plattnerite after 24 hours of reaction at pH 7.6 with
50 mg C/L DIC and different iodide concentrations .............................................56
4.10
Electron micrographs of solids: (a) before reaction, and after 24 hours of reaction in
the CSTR with 50 mg C/L DIC and (b) 10 μM iodide at pH 7.6, and (c) 100
μM iodide at pH 7.6 ...............................................................................................57
4.11
Dissolution rate of plattnerite at pH 7.6 in the absence and presence of 50 mg C/L DIC
with different iodide concentrations ......................................................................58
4.12
Effect of DIC on the dissolution rate of plattnerite with 10 μM iodide at pH 7.6 and
8.5...........................................................................................................................59
4.13
Pathways of the reductive dissolution of PbO2 by iodide ..................................................61
4.14
Model for PbO2 dissolution rate in the presence of iodide at pH 7.6 and 8.5....................62
4.15
The dissolution rate of plattnerite with 50 mg C/L DIC in the absence and presence of
1 mg Cl2/L free chlorine at pH 5.7 to 8.5...............................................................63
4.16
Effect of free chlorine concentration on the dissolution rate of plattnerite with 50 mg
C/L DIC at pH 7.6 and 8.5 .....................................................................................64
4.17
Effect of orthophosphate concentration on the dissolution rate of plattnerite with 0
and 50 mg C/L DIC at pH 7.6 and 8.5 ...................................................................65
4.18
Conceptual model of the potential pathways of PbO2 dissolution in the presence of
orthophosphate .......................................................................................................65
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
xii | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
4.19
Electron micrographs of (a) plattnerite before reaction, (b) pure hydroxylpyromorphite,
(c) solids after reaction of plattnerite at pH 7.5, 0 mg C/L DIC, 1 mg P/L
phosphate, and 2 mg Cl2/L monochloramine, and (d) solids after reaction of
plattnerite at pH 8.5, 50 mg C/L DIC, 1 mg P/L orthophosphate, and 2 mg
Cl2/L .......................................................................................................................66
4.20
Steady-state concentrations for plattnerite dissolution in the presence of 1 mg P/L
orthophosphate and predicted lead concentration in equilibrium with
hydroxylpyromorphite ...........................................................................................67
B.1
X-ray diffraction patterns of solids following reaction of massicot at pH 10 in the
absence of DIC with 20 mg Cl2/L free chlorine ....................................................77
B.2
X-ray diffraction patterns of solids following 28 days of reaction of massicot with 20
mg C/L DIC and 4 mg/L as Cl2 free chlorine ........................................................78
B.3
X-ray diffraction patterns of solids following reaction of massicot at pH 10 in the
presence of 20 mg C/L DIC with 20 mg Cl2/L free chlorine .................................79
B.4
X-ray diffraction patterns of solids following reaction of massicot at pH 7.5 in the
presence of 20 mg C/L DIC with 20 mg Cl2/L free chlorine .................................80
B.5
X-ray diffraction patterns of solids following reaction of hydrocerussite at pH 7.5 in
the absence DIC with 20 mg Cl2/L free chlorine ...................................................81
B.6
X-ray diffraction patterns of solids following reaction of hydrocerussite at pH 7.5 in
the absence DIC with 20 mg Cl2/L free chlorine ...................................................82
B.7
X-ray diffraction patterns of solids following reaction of cerussite at pH 7.5 in the
absence DIC with 20 mg Cl2/L free chlorine .........................................................83
B.8
X-ray diffraction patterns of solids following reaction of cerussite at pH 10 in the
absence DIC with 20 mg Cl2/L free chlorine .........................................................84
C.1
Dissolved lead concentration from batch reactors over time at pH 7.6 with 50 mg C/L
DIC and different concentrations of MOPS...........................................................86
C.2
Effluent lead concentrations from CSTRs as a function of the number of hydraulic
residence times (τ = 30 min) at pH 7.6 with 50 mg C/L DIC in the absence
and presence of 1 mM MOPS ................................................................................87
C.3
Effluent lead concentrations and pH from CSTRs as a function of the number of
hydraulic residence times (τ = 30 min) at pH 7.6 with different iodide
concentrations. .......................................................................................................88
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
FOREWORD
The Water Research Foundation (Foundation) is a nonprofit corporation dedicated to the
development and implementation of scientifically sound research designed to help drinking
water utilities respond to regulatory requirements and address high-priority concerns. The
Foundation’s research agenda is developed through a process of consultation with Foundation
subscribers and other drinking water professionals. The Foundation’s Board of Trustees and
other professional volunteers help prioritize and select research projects for funding based upon
current and future industry needs, applicability, and past work. The Foundation sponsors
research projects through the Focus Area, Emerging Opportunities, and Tailored Collaboration
programs, as well as various joint research efforts with organizations such as the U.S.
Environmental Protection Agency and the U.S. Bureau of Reclamation.
This publication is a result of a research project fully funded or funded in part by
Foundation subscribers. The Foundation’s subscription program provides a cost-effective and
collaborative method for funding research in the public interest. The research investment that
underpins this report will intrinsically increase in value as the findings are applied in
communities throughout the world. Foundation research projects are managed closely from their
inception to the final report by the staff and a large cadre of volunteers who willingly contribute
their time and expertise. The Foundation provides planning, management, and technical
oversight and awards contracts to other institutions such as water utilities, universities, and
engineering firms to conduct the research.
A broad spectrum of water supply issues is addressed by the Foundation's research
agenda, including resources, treatment and operations, distribution and storage, water quality and
analysis, toxicology, economics, and management. The ultimate purpose of the coordinated
effort is to assist water suppliers to provide a reliable supply of safe and affordable drinking
water to consumers. The true benefits of the Foundation’s research are realized when the results
are implemented at the utility level. The Foundation's staff and Board of Trustees are pleased to
offer this publication as a contribution toward that end.
Roy L. Wolfe, Ph.D.
Chair, Board of Trustees
Water Research Foundation
Robert C. Renner, P.E.
Executive Director
Water Research Foundation
xiii
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
ACKNOWLEDGMENTS
The authors appreciate the advice and guidance of the Project Advisory Committee
(PAC) – France Lemieux, Health Canada and Becki Rosenfeldt, Hazen and Sawyer - and the
Water Research Foundation Project Manager, Traci Case. Windsor Sung of the Massachusetts
Water Resources Authority also provided helpful input over the course of this project. The
authors acknowledge the assistance of Wenlu Li, Zimeng Wang, and Jiewei Wu in conducting
selected experiments and associated analyses.
xv
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
EXECUTIVE SUMMARY
OBJECTIVES
The primary objective of this project was to advance our understanding of the dynamic
processes that form PbO2 solids in drinking water distribution systems and that control their
subsequent stability and dissolution. Three specific objectives were pursued.
1. Identify threshold values of water chemistry parameters that govern the formation of
PbO2 and control which PbO2 solid (plattnerite vs. scrutinyite) is produced.
2. Establish the equilibrium solubility of PbO2 solids at conditions for which lead remains in
the +IV oxidation state.
3. Elucidate the mechanisms and quantify the rates of the coupled processes of PbO2
reduction and dissolution.
BACKGROUND
Lead release from pipes in distribution systems is a serious threat to public health. The
observations of extremely high lead levels in Washington D.C. tap water in 2000-2004 illustrate
the importance of this problem (Edwards and Dudi 2004; Renner 2004; U.S. Environmental
Protection Agency 2007b). Research can contribute to the scientific basis for corrosion control
strategies and help the water supply community meet the challenge of an aging infrastructure.
Lead concentrations in drinking water are affected by chemical reactions that occur
within the distribution system. Of particular relevance to management of lead concentrations are
the reactions that control the formation and stability of lead(IV) oxides (PbO2 solid phases of
scrutinyite and plattnerite). PbO2 solids form at the high oxidation-reduction potential induced
by residual free chlorine, and such solids have been observed as constituents of scales of lead
corrosion products that develop on lead pipes (Schock et al. 2001; Edwards and Dudi 2004; Lytle
and Schock 2005). The PbO2 solids have low solubility. As long as a sufficiently high
oxidation-reduction potential is maintained, dissolved lead concentrations remain at low levels.
However, when the oxidation-reduction potential is lowered, as can occur when switching to
chloramine as a residual disinfectant, the PbO2 is no longer stable and its reduction releases lead
to the water. Even when oxidizing conditions are present, the actual solubility of PbO2 phases is
imprecisely known.
Information on the dissolution rates of PbO2 is particularly valuable as water suppliers
consider process changes that affect water chemistry such as switching disinfectant type or dose,
adjusting pH, or adding a corrosion inhibitor. Recent research linked the high lead
concentrations in the Washington D.C. service area from 2000-2004 to the breakdown of PbO2
solids following a change in the residual disinfectant from free chlorine to chloramine (Edwards
and Dudi 2004; Renner 2004; U.S. Environmental Protection Agency 2007b). For other utilities
that are considering switching from free chlorine to chloramine, information is needed on the
rates and mechanisms of PbO2 dissolution as a function of water chemistry. Orthophosphate has
been demonstrated to decrease dissolved lead concentrations for systems using chloramine
(Edwards and McNeill 2002; U.S. Environmental Protection Agency 2007b). Strategies that
xvii
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
xviii | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
inhibit PbO2 dissolution and minimize lead release can be optimized based on an understanding
of the effects of water chemistry on lead release rates.
APPROACH
The project was divided into three integrated tasks that corresponded to the three research
objectives (Figure ES.1). The tasks progressed from (1) the formation of PbO2, to (2) the
stability of PbO2 in equilibrated systems, and finally to (3) the rates of dissolution of PbO2. All
tasks integrated dissolved phase analysis with characterization of the structures and compositions
of the solid phases; this approach facilitated observation of macroscopic precipitation and
dissolution processes while developing a mechanistic basis for those processes.
Figure ES.1 Integrated approach to investigate dynamics of PbO2 in distribution
Task 1: Critical Water Chemistry Thresholds for the Formation of PbO2
The impact of water chemistry on the extent of PbO2 formation and the identity of the
PbO2 phases formed was evaluated in a series of bench-scale laboratory experiments. The
formation of PbO2 by the oxidation of different lead(II) and lead(0) precursors in water was
investigated as a function of free chlorine concentration, pH, dissolved inorganic carbon (DIC),
and time. Experiments were performed with five different precursor materials: elemental lead(0)
powder, dissolved lead(II) chloride (PbCl2), the lead(II) oxide massicot (β-PbO), the lead(II)
carbonate cerussite (PbCO3), and the lead(II) hydroxycarbonate hydrocerussite (Pb(OH)2(CO3)2).
Aqueous and solid phase samples were collected over reaction times up to 28 days. The
presence of PbO2 in the products of the reaction was identified using scanning electron
microscopy (SEM) and X-ray diffraction (XRD). These methods were also used to identify
intermediate products of the reaction that could influence the rate of PbO2 formation and the
identity of the specific PbO2 solid that formed (scrutinyite versus plattnerite).
Task 2: Determination of Equilibrium Solubility of PbO2
The dissolved lead concentrations achieved in aqueous solutions equilibrated with the
two PbO2 polymorphs, scrutinyite and plattnerite, were measured as a function of pH at free
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Executive Summary | xix
chlorine concentrations designed to keep lead in the +IV oxidation state. Batch experiments
were conducted with pure forms of plattnerite and scrutinyite. Filtered samples were collected
over time to determine the concentration at which dissolved lead stabilized and the time to reach
that stable state. Solids were also collected and characterized by XRD and SEM to probe for any
transformations of the solids that might have occurred during the equilibrations.
Task 3: Coupled Reduction-Detachment Processes During Dissolution of PbO2
The dissolution rates of plattnerite were examined as a function of water chemistry
parameters that included pH, DIC, iodide, free chlorine, and orthophosphate. Because PbO2 is
often found in systems that are not at equilibrium, information about rates and not equilibrium
solubility may be more important for predicting lead concentrations in water that has been in
contact with corrosion scales that contain PbO2. Dissolution rates were quantified using
continuously-stirred tank reactors loaded with plattnerite and fed influents of controlled
composition. Four sets of experiments probed different impacts of water chemistry on the
plattnerite dissolution rate.
 Baseline experiments with DIC over a range of pH to determine the rates in the absence
of any oxidants, reductants, or corrosion inhibitors.
 Experiments with iodide as a chemical reductant and carbonate from DIC as a Pb(II)complexing ligand to probe the mechanisms of the reductive dissolution of PbO2 as a
process involving chemical reduction and detachment.
 Experiments with free chlorine present to quantify its inhibitory effect on PbO2
dissolution and to probe for a critical concentration required to limit lead release.
 Experiments with orthophosphate present as a potential corrosion inhibitor.
RESULTS AND CONCLUSIONS
The research project made significant findings regarding the formation and stability of
lead(IV) oxides at conditions relevant to drinking water distribution systems. Key findings with
respect to formation, equilibrium dissolved lead concentrations, and PbO2 dissolution rates are
outlined below.
Formation of Lead(IV) Oxides



PbO2 can only form in the presence of free chlorine, and the threshold free chlorine
concentration for producing PbO2 is less than 4 mg Cl2/L.
The formation of PbO2 is accelerated by the presence of dissolved inorganic carbon.
Both polymorphs of PbO2 (plattnerite and scrutinyite) can form, and most conditions
resulted in a mixture of the two phases. The identity of the Pb(II)-containing precursor
could control the identity of the specific PbO2 phase that formed.
Dissolved Lead Concentrations in Equilibrated Solutions

Dissolved lead concentrations from equilibration of PbO2 in water with free chlorine
were orders of magnitude higher than predicted from published thermodynamic data.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
xx | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems



Low dissolved lead concentrations in chlorinated solutions were still maintained over
long periods of time. It was only after long stagnation times of 4 days or longer that the
action level was reached at pH 7.5 and 8.5, and the action level was never reached at pH
6.0.
The final dissolved lead concentrations achieved upon long-term contact of plattnerite
and scrutinyite with water were similar.
Dissolved lead concentrations increased with increasing pH above pH 6 and were
unaffected by the presence of DIC.
Dissolution Rates of PbO2





The dissolution rate and not the equilibrium solubility of PbO2 will control dissolved lead
concentrations in waters that are in contact with PbO2 as a corrosion product for most
relevant stagnation times.
The rate of PbO2 dissolution rate decreased with increasing pH. This is in contrast to the
dissolved lead concentrations after multi-day equilibration that increased with increasing
pH.
Dissolution was strongly inhibited by free chlorine even at concentrations as low as 0.2
mg Cl2/L.
Orthophosphate inhibited PbO2 dissolution with its effects limited to near-neutral pH.
The presence of chemical reductants can significantly accelerate the dissolution of PbO2.
Dissolution of PbO2 is a two-step process that involves chemical reduction of Pb(IV) to
Pb(II) followed by detachment of Pb(II) and release to solution. For most conditions the
chemical reduction steps will be rate-limiting.
APPLICATIONS AND RECOMMENDATIONS
Lead(IV) oxides (PbO2) can be an important component of corrosion products on pipe
scales for utilities that have lead service lines in their distribution systems and that currently use
or have used free chlorine as the secondary disinfectant. PbO2 can only form in systems with
free chlorine present; however, because of the low solubility of this phase and the complexity of
pipe scales, PbO2 may persist well after a switch from free chlorine to chloramine. The rate of
PbO2 formation and consequently the likely extent of PbO2 formation on lead service lines is
strongly affected by the water chemistry of the distribution system. The susceptibility of PbO2 to
reductive dissolution and its likely enrichment at the surface of the pipe scale can make it a
major contributor to lead release to water even when it is only present as a minor constituent of a
pipe scale. The presence of PbO2 is a more important finding then the exact identity of the form
of PbO2 (plattnerite versus scrutinyite) that is present. Both plattnerite and scrutinyite can form
on lead pipes in the presence of free chlorine, but similar dissolved lead concentrations were
achieved in aqueous solutions equilibrated with the two solids. When free chlorine is present,
low dissolved lead concentrations below the action level can be achieved even for stagnation
times of several days.
For PbO2-containing systems without free chlorine or for which free chlorine has been
depleted, dissolved lead concentrations will be controlled by the rate of the dissolution reaction
and not by equilibrium solubility. The dissolution rate of PbO2 is a very strong function of the
water chemistry, and orders of magnitude differences can occur in the rates depending on the
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Executive Summary | xxi
composition. Utilities should be aware of conditions that could accelerate the dissolution of
PbO2 and the release of lead from this potential reservoir of unstable lead in scales on lead
service lines. This is particularly important for systems that have recently switched from using
free chlorine to chloramine as the secondary disinfectant. Dissolution rates of PbO2 will increase
in response to process changes that lower the pH of the water in the distribution system. The
most significant parameter affecting PbO2 dissolution rates is the concentration and identity of
species that can act as chemical reductants to accelerate PbO2 dissolution. These species include
natural organic matter, dissolved iron(II) and manganese(II), and iodide. The impacts of water
chemistry on the dissolution rate of PbO2 can also be used to promote conditions with the
slowest dissolution rates and consequently the best abilities to maintain low lead concentrations
in distribution systems. Free chlorine is an excellent inhibitor of the reductive dissolution of
PbO2 even at concentrations as low as 0.2 mg/L as Cl2. Orthophosphate can also be effective at
limiting the release of lead to water in systems that contain PbO2 for systems with pH values
near neutral.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
CHAPTER 1
INTRODUCTION
SIGNIFICANCE OF LEAD(IV) OXIDES TO LEAD IN DRINKING WATER
Lead release from pipes in distribution systems is a serious threat to public health. The
observations of extremely high lead levels in Washington D.C. tap water early in the last decade
illustrate the importance of this problem (Edwards and Dudi 2004; Renner 2004; U.S.
Environmental Protection Agency 2007b). While advances in water treatment, supply, and
distribution were ranked fourth on the list of 20th Century engineering achievements compiled by
the National Academy of Engineering, management of our water treatment and supply
infrastructure has been identified as a great engineering challenge for the 21st century, and
corrosion control plays a leading role in this challenge (Edwards 2004). The American Society
of Civil Engineers gave drinking water a D- in its 2004 report card for infrastructure (American
Society of Civil Engineers 2005). As the water supply community meets the challenge of an
aging infrastructure, there is an increasing need for research to provide a scientific basis for
corrosion control.
Lead concentrations in drinking water are affected by chemical reactions that occur
within the distribution system. Of particular relevance to management of lead concentrations are
the reactions that control the formation and stability of lead(IV) oxides (PbO2 solid phases of
scrutinyite and plattnerite). PbO2 solids form at the high oxidation-reduction potential induced
by residual free chlorine, and such solids have been observed as constituents of scales of lead
corrosion products that develop on lead pipes (Edwards and Dudi 2004; Lytle and Schock 2005).
The PbO2 solids have low solubility, and as long as a sufficiently high oxidation-reduction
potential is maintained, dissolved lead concentrations remain at low levels. However, when the
oxidation-reduction potential is lowered, as can occur when switching to chloramine as a residual
disinfectant, the PbO2 is no longer stable and its reduction releases lead to the water. Even when
highly oxidizing conditions are present, the actual solubility of PbO2 phases is imprecisely
known.
Information on the dissolution rates of PbO2 is particularly valuable as water suppliers
consider process changes that affect water chemistry such as switching disinfectant type or dose,
adjusting pH, or adding a corrosion inhibitor. Recent research linked the high lead
concentrations in the Washington D.C. service area from 2000-2004 to the breakdown of PbO2
solids following a change in the residual disinfectant from free chlorine to chloramine (Edwards
and Dudi 2004; Renner 2004; U.S. Environmental Protection Agency 2007b). For other utilities
that are considering switching from free chlorine to chloramine, information is needed on the
rates and mechanisms of PbO2 dissolution as a function of water chemistry. Orthophosphate has
been demonstrated to decrease dissolved lead concentrations for systems using chloramine
(Edwards and McNeill 2002; U.S. Environmental Protection Agency 2007b). Decreases occur
through the precipitation of low solubility Pb(II) phosphate solids, but the identity of these solids
and the rates and mechanisms of their formation have not been established. This research project
examined the mechanisms through which PbO2 dissolves and the potential for orthophosphate
and low levels of free chlorine to mitigate lead release. An improved understanding of the
effects of water chemistry on lead release rates can be used to identify strategies that inhibit
PbO2 dissolution and minimize lead release when dissolution occurs.
1
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
2 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
PROJECT OBJECTIVES, HYPOTHESES, AND STRUCTURE OF THE REPORT
Project Objectives
The project objective is to advance our understanding of the dynamic processes that form
PbO2 solids in drinking water distribution systems and that control their subsequent stability and
dissolution. Three specific objectives and associated hypotheses are described below. The
research approach is organized into three tasks to achieve these objectives.
Objective 1: Identify threshold values of water chemistry parameters that govern the
formation of PbO2 and control which PbO2 solid (plattnerite vs. scrutinyite) is produced.
Hypothesis 1A: A threshold concentration of free chlorine must be exceeded for the
formation of PbO2, and the presence of dissolved inorganic carbon will favor the formation of
Pb(II) carbonate solids that are intermediate phases in PbO2 formation.
Hypothesis 1B: Plattnerite and scrutinyite will both form upon reaction of Pb(0) and
Pb(II) solids with free chlorine, and the specific phase formed will be governed by the initial
nucleation rate and the Pb(0) or Pb(II) precursor.
Significance: Many drinking water utilities use a free chlorine disinfectant residual, but
not all of their distribution systems may contain PbO2 solids. Knowledge of the extent of PbO2
formation and the identity of the PbO2 phases present is essential to predicting the potential for
lead release upon changes to the water chemistry.
Objective 2: Establish the equilibrium solubility of PbO2 solids at conditions for which lead
remains in the +IV oxidation state.
Hypothesis 2A: The equilibrium solubility of PbO2 at highly oxidizing conditions will
increase with decreasing pH and will be below the drinking water standard for the relevant pH
range.
Hypothesis 2B: The equilibrium solubility of scrutinyite and plattnerite will be similar
(within an order of magnitude).
Significance: In distribution systems with PbO2 and a sustained free chlorine residual,
the dissolved lead concentrations will be controlled by the solubility of this phase. Measurement
of the equilibrium solubility of PbO2 will fill an important gap in our ability to predict dissolved
lead concentrations in distribution systems.
Objective 3: Elucidate the mechanisms and quantify the rates of the coupled processes of
PbO2 reduction and dissolution.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 1: Introduction | 3
Hypothesis 3A: The net release of lead from PbO2 solids will be governed by coupled
steps of (1) Pb(IV) reduction to Pb(II) and (2) detachment of Pb(II) from the surface. The water
chemistry will govern which of these two steps is rate-limiting.
Hypothesis 3B: Phosphate will effectively mitigate lead release from PbO2 by
sequestering Pb(II) in low solubility Pb(II) phosphate solid phases in the pyromorphite group of
minerals.
Significance: The breakdown of PbO2 present in pipe scales has caused elevated lead
concentrations. Knowledge of the specific processes responsible for PbO2 breakdown and the
associated rates of those processes can be used to design mitigation strategies. The addition of
phosphate has been demonstrated to limit Pb release by forming Pb(II) phosphates, but the rates
and mechanisms of their formation are not currently known.
Structure of the Report
The project was organized into three tasks (Figure 1.1). Each task corresponds to one of
the three research objectives just presented. All tasks integrated dissolved phase analysis with
characterization of the structures and compositions of the solid phases; this approach facilitated
observation of macroscopic precipitation and dissolution processes while developing a
mechanistic basis for those processes. In Task 1 we determined the impact of water chemistry
on the extent of PbO2 formation and the identity of the PbO2 phases formed. The formation of
PbO2 from different Pb(II) precursors was investigated as a function of pH, time, and the
concentrations of free chlorine and dissolved inorganic carbon. The results of Task 1 are
presented in Chapter 2 of this report. Chapter 3 reports the results of Task 2, a task in which we
quantified the dissolved lead concentrations that are maintained when chlorinated water
equilibrated with scrutinyite or plattnerite. Dissolved lead concentrations were measured as a
function of pH at free chlorine concentrations that help to keep lead in the Pb(IV) oxidation state.
In Task 3, presented in Chapter 4, we determined the rate of lead release from PbO2 when
dissolution occurs by coupled processes of Pb(IV) reduction and detachment of Pb(II) from the
solid surface. A series of experiments investigated the influence of water chemistry on the rates
Figure 1.1 Integrated approach to investigate dynamics of PbO2 in distribution systems.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
4 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
of dissolution and the mitigation of lead release.
Following the presentation of the results from the three tasks in Chapters 2-4, a research
summary and conclusions are included as Chapter 5. Chapter 6 then presents a summary of
recommendations to utilities based on the findings of this project.
BACKGROUND ON LEAD(IV) OXIDES AND RELEVANCE TO DRINKING WATER
Lead is a toxic heavy metal and the adverse effects of lead consumption are a public
health concern. While water leaving treatment plants has very low concentrations of lead,
internal corrosion of lead-containing pipe, fittings, and solder in water distribution systems can
contribute lead to drinking water. While new construction does not use lead pipe, many older
buildings retain the original lead service lines and internal plumbing. Concern for lead
concentrations in water distribution systems motivated the passage of the Lead and Copper Rule
in 1991. Controlling lead concentrations can be particularly challenging for drinking water
utilities with older distribution system infrastructure and source waters with neutral pH and low
alkalinity. The recent observations of high lead levels in Washington D.C. tap water
demonstrate the magnitude of this challenge (Edwards and Dudi 2004; Renner 2004), and other
North American cities face similar challenges (Nour et al. 2007; Renner 2006; 2007).
Chemical reactions that occur within the distribution system govern the dissolved lead
concentrations in drinking water. Lead release can occur within components owned by the
utility or in the plumbing of individual connections. Developing a better understanding of
responses of premise plumbing to water quality changes was identified as one of three primary
goals in the Water Research Foundation’s 2007 Distribution System Water Quality Strategic
Initiative Plan (Awwa Research Foundation 2007). The release of lead from metal pipes and
fittings is controlled by both the water composition and the properties of the pipe surface. Lead
may be released directly from the pipe or from lead-containing corrosion products on the pipe
surface. Corrosion products that develop in scales on pipe surfaces include lead(II) carbonates,
phosphates, and oxides (Schock and Lytle 2010). More recently lead(IV) oxides (PbO2), which
are the focus of this project, have been recognized as important components in pipe scales (Lytle
and Schock 2005; Schock et al. 2008).
Information on the formation and stability of PbO2 is important as water suppliers
consider process changes that affect water chemistry (AWWA 2005). Knowledge of the
conditions favorable for the development of PbO2 can provide estimates of the extent and
identity of these solid phases within a distribution system. As long as the water chemistry
remains sufficiently oxidizing (e.g., from a sufficient free chlorine residual), dissolved lead
concentrations in equilibrium with PbO2 will be kept at low levels. However, changes in water
chemistry that lower the oxidation-reduction potential can destabilize the PbO2 solids and cause
release of lead to the water. In addition to releasing lead, the breakdown of lead corrosion
products can also mobilize other contaminants that are enriched in the corrosion products
(Schock et al. 2008).
The increase in lead concentrations in the Washington D.C. Water and Sewer Authority
service area in 2000-2004 provide a dramatic example of the impacts of water chemistry on
PbO2 formation and dissolution. Lead(IV) oxides developed on pipe scales when free chlorine
was used as the residual disinfectant.
When residual disinfection was switched to
chloramination, lead release from the pipe resulted in high lead concentrations (Edwards and
Dudi 2004; Renner 2004; U.S. Environmental Protection Agency 2007b).
The lead
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 1: Introduction | 5
concentrations have ultimately been controlled by increasing the pH and adding orthophosphate
(Edwards and McNeill 2002; U.S. Environmental Protection Agency 2007b), which further
illustrates the influence of water chemistry on rates of lead release from pipe surfaces.
Lead(IV) Oxide Occurrence and Stability in Distribution Systems
Occurrence of Lead(IV) Oxides
Lead(IV) oxides have been observed as a component of the scales on lead pipe from
distribution systems (Lytle and Schock 2005; Schock et al. 2008; Vasquez et al. 2006). Two
polymorphs of PbO2 can be formed: scrutinyite (α-PbO2) and plattnerite (β-PbO2) (Lytle and
Schock 2005). In a study of pipe sections removed from 34 different distribution systems that
represented a range of water chemistries, 26% of the systems had samples that included PbO2.
Investigation of these pipe scales as a function of depth observed three different modes of PbO2
occurrence:
uniform scales, patchy coverage with coexistence with hydrocerussite
(Pb3(CO3)2(OH)2), and a layered structure with plattnerite (β-PbO2) on top of cerussite (PbCO3)
(Schock et al. 2005b). In a laboratory study that contacted new lead pipe with chlorinated water
(3.5 mg/L as Cl2) at a relatively high pH of 9.5-10, both plattnerite and scrutinyite formed in
scales of corrosion products on the pipe surface. They only formed after longer overall contact
times than were necessary for the formation of lead(II) carbonate corrosion products, and even
when the PbO2 solids were observed, they were present together with lead(II) carbonates(Xie and
Giammar 2011).
Although lead is not stable in the +IV oxidation state in oxygenated water, this oxidation
state is stable at the high oxidation-reduction potential (ORP) provided by free chlorine. The
formation of PbO2 was observed upon reaction of dissolved lead with sodium hypochlorite, and
Pb(II) carbonate phases were observed as intermediate products that can co-exist with PbO2.
The formation of PbO2 is reversible, with PbO2 disappearing when a residual hypochlorite
concentration is not maintained. Hydrocerussite can be transformed to scrutinyite following a
lag period when reacted with free chlorine (Liu et al. 2006). In contrast to the facilitated
formation of PbO2 by carbonate, the presence of orthophosphate inhibited PbO2 formation in
chlorinated solutions (Lytle et al. 2009).
Solubility of Lead(IV) Oxides
In the presence of strong oxidants, HOCl/OCl- in particular, lead can be oxidized from
the +II oxidation state to the +IV oxidation state. Figure 1.2 illustrates the dominant solid and
dissolved species as a function of both pH and ORP. The reduction half-reaction is given as
Equation 1.1.
Pb4+ + 2e- = Pb2+,EH° = 0.845 V
(1.1)
A complete reaction for the oxidation of dissolved lead(II) by HOCl to form PbO2(s) is given as
Equation 1.2.
Pb2+ + HOCl + H2O = PbO2(s) + Cl- + 3H+
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
(1.2)
6 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Figure 1.2 Predominance areas of lead(0)/lead(II)/lead(IV) system for 15 μg/L total
lead and 30 mg/L dissolved inorganic carbon.
a)
-5
b)
-7
-1
-2
-9
Log[Pb]diss (M)
Log[Pb]diss (M)
-3
-11
-13
equilibrium
w ith PbO2
-15
-17
-19
-4
-5
equilibrium w ith
hydrocerussite
-6
-7
-21
equilibrium w ith
hydroxylpyromorphite
-8
-23
-25
-9
4
5
6
7
8
pH
9
10
11
4
5
6
7
8
9
10
11
pH
Figure 1.3 Dissolved lead concentration in equilibrium with (a) lead(IV) oxide at
sufficiently oxidizing conditions that all lead is present as lead(IV) and (b) the lead(II)
carbonate hydrocerussite with 30 mg/L DIC and the lead(II) phosphate
hydroxylpyromorphite with 30 mg/L DIC and 3 mg/L orthophosphate. Note the
difference in the y-axis ranges.
For this case, the divalent free lead ion concentration is determined by the oxidationreduction potential (ORP) set by the concentrations of chloride and hypochlorous acid and by the
pH. When lead(IV) is the dominant oxidation state, as controlled by free chlorine, the dissolved
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 1: Introduction | 7
lead concentrations can be very low (Figure 1.3a). The lead(IV) oxide (PbO2) that forms has a
very low solubility given by the reaction and solubility product in Equations 1.3 and 1.4.
PbO2(s) + 4H+ = Pb4+ + 2H2O
K sp , scrutinyite 
4
[ Pb ]
 10 8.26
 4
[H ]
(1.3)
(1.4)
The total dissolved lead(IV) is the sum of the concentrations of the free metal ion Pb4+ as well as
the hydrolysis complexes PbO32- and PbO44-; however, these complexes are only significant at
very high pH (> 12) and are negligible over the pH range of interest in water supply and
treatment (Pourbaix 1974). While free chlorine is sufficiently oxidizing to promote the
formation of lead(IV), monochloramine can oxidize metallic lead(0) to lead(II) but not to
lead(IV) (Switzer et al. 2006). Direct electrochemical measurement of the ORP values of
monochloramine and free chlorine confirmed the inability of monochloramine to stablilize
lead(IV) oxides (Rajasekharan et al. 2007).
Stability of Lead(IV) Oxides
If a strong oxidant like HOCl/OCl- is no longer present, then the lead(IV) in PbO2 can be
reduced to lead(II). The dissolved lead concentrations in equilibrium with lead(II) carbonates
and phosphates are significantly higher than those in equilibrium with lead(IV) oxides (Figure
1.3b). While recent studies have established the ability of natural organic matter (Dryer and
Korshin 2007; Lin and Valentine 2008a; Lin and Valentine 2009; Shi and Stone 2009b), iodide
(Lin et al. 2008), bromide at acidic conditions (Lin and Valentine 2010), and iron(II) and
manganese(II) (Shi and Stone 2009a) to reduce PbO2. These studies and others (Switzer et al.
2006) have also shown that PbO2 can undergo reduction even in solutions without strong
reductants. PbO2 is such a strong oxidant, that its reduction by water is even energetically
favorable (Equation 1.5).
PbO2(s) + 2H+ = Pb2+ + 0.5O2 + H2O
(1.5)
Maintaining the low solubility of PbO2 requires that the water chemistry be sufficiently
oxidizing to maintain lead in the +IV oxidation state. This has been observed by the requirement
of free chlorine in contact with lead pipes to provide a low dissolved lead concentration (Boyd et
al. 2007). Changes in ORP that result from changes in the residual disinfectant or in disinfectant
demand in the distribution system can lead to transformations from low solubility PbO2(s) to
higher solubility lead(II) carbonate and phosphate phases (Schock and Giani 2004). A study of
lead release from pipes that included PbO2 as a constituent of the pipe scale observed lead
concentrations below 20 µg L-1 in chlorinated water held stagnant in the pipes until the free
chlorine was depleted, and after that depletion the lead concentrations increased markedly (Xie
and Giammar 2011). The decrease in the ORP that occurs when chloramine is used instead of
free chlorine causes the PbO2 to breakdown and release lead back to solution (Vasquez et al.
2006). Laboratory and pilot-scale studies have observed higher dissolved lead levels with
chloramine than with free chlorine, which can be explained by the differences in solubility of
Pb(IV) and Pb(II) phases (Edwards and Dudi 2004; Vasquez et al. 2006).
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
8 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Dissolution Rates of Lead Corrosion Products
Information on the dissolution rates of lead-containing solids is necessary to complement
knowledge of the equilibrium solubility. The water present in a drinking water distribution will
often not be in equilibrium with the solid phases present in the pipe. Estimates of lead
concentrations in distribution systems based on equilibrium solubility are qualitatively good, but
they are usually not quantitatively accurate. Equilibrium-based models tend to overpredict lead
concentrations. Such models are limited by the accuracy of available equilibrium constants,
transitions between scale types, by-product release, and reaction kinetics (Edwards et al. 1999;
Vasquez et al. 2006). For systems without equilibrium between the solid and dissolved phase,
the rate of dissolution and the contact time between the water and the lead-containing solid will
determine the dissolved lead concentration. Differences between lead concentrations during
times of flowing water and during stagnation periods can be significant. For example, lead
concentrations in new pipe were low during flowing conditions but exceeded 15 μg/L after 8
hours of stagnation, even in the presence of inhibitors (Hozalski et al. 2005). During stagnant
periods, dissolved lead concentrations in lead pipe approach equilibrium exponentially with time.
The largest increases occur within the first 24 hours (Lytle and Schock 2000).
Dissolution rates of precipitated solids are functions of properties of both the dissolving
solid and the solution in which it is dissolving. Dissolution rates generally increase with
decreasing pH, with this effect becoming most pronounced at neutral pH and below. Rates can
be affected by inhibiting or enhancing dissolved species that include orthophosphate, carbonate,
other complexing ligands, and reductants. Reductants will accelerate the dissolution of PbO2.
The rapid release of lead from distribution systems following switches from free chlorine to
chloramine indicates the large energetic driving force for reductive dissolution of PbO2.
The dissolution rates of lead(II) carbonate and phosphate solids have been examined in
previous research as a function of water chemistry, in particular as a function of pH. However,
very little was known about the rates of PbO2 dissolution and the factors controlling those rates.
With previous WaterRF support, we prepared a thorough review of the literature of dissolution
rates of lead corrosion products and measured the dissolution rates of different lead(II) solids
(Giammar et al. 2010). Carbonate solids are among the most rapidly soluble minerals (Stumm
and Morgan 1996), and their dissolution rates increase significantly with decreasing pH at acidic
conditions. Dissolution rates of PbCO3(s) were on the order of 10-4 mol m-2 h-1 at pH 7
(Pokrovsky and Schott 2002), and our recent work measured rates of 1.2·10-6 mol m-2 h-1 at pH
7.5 in water with no added alkalinity, orthophosphate, or disinfectant (Giammar et al. 2010). For
lead(II) phosphates, a recent study on the dissolution rates of chloropyromorphite
(Pb5(PO4)3Cl(s)) observed that dissolution rates increase with decreasing pH and decreasing
solution saturation; at a pH of 7 and a highly undersaturated solution, the dissolution rate is
1.1·10-7 mol m-2 h-1 (Xie and Giammar 2007), and for hydroxylpyromorphite a dissolution rate of
1.2·10-8 mol m-2 h-1 was measured at pH 7.5 in a simple aqueous solution (Giammar et al. 2010).
Only recently have dissolution rate measurements of lead(IV) oxides been made.
Because the reductive dissolution of PbO2 in pure water is energetically favorable, extensive
dissolution was measured for plattnerite (Lin and Valentine 2008a) and scrutinyite (Dryer and
Korshin 2007). The dissolution of both materials was significantly accelerated by the presence
of natural organic matter (NOM), which served as a reductant for the PbO2. NOM may also
inhibit the development of Pb(II) precipitates that would limit further dissolution. The
dissolution of PbO2 increased with decreasing pH. At pH 7, the data of Lin and Valentine
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 1: Introduction | 9
(2008a) can be used to extract a dissolution rate of 6.5·10-9 mol m-2 h-1. This study also verified
that the lead released was lead(II) and not lead(IV).
Mitigation of Lead Release from Corrosion Products on Lead Pipe
Many water utilities must implement a corrosion control strategy to minimize dissolved
lead concentrations in their distribution systems. Recommended strategies for controlling lead
corrosion include pH and/or alkalinity adjustment and addition of orthophosphate (Edwards et al.
1999). The general strategy of corrosion control is to promote the formation of low solubility
lead-containing solids that passivate the surface. These solids include lead(IV) oxides at
sufficiently oxidizing conditions; however, when lead(IV) oxides are no longer stable, then the
impact of water chemistry on the stability of lead(II) phases becomes important.
Dissolved lead concentrations can respond significantly to changes in pH and alkalinity.
Based on a survey of utilities, large decreases in soluble lead levels are observed with increasing
alkalinity (Edwards et al. 1999). The effect of increasing alkalinity is consistent with lead
concentrations being determined by the solubility of lead carbonate (cerussite and
hydrocerussite) solids. The stability of these carbonate solids is determined by the activity of the
carbonate ion, which is determined by both the DIC (or alternatively alkalinity) and pH.
Consequently, increasing pH without also increasing carbonate alkalinity will not be as effective.
In fact, without sufficient alkalinity (less than 20-30 mg/L as CaCO3), increasing pH can actually
exacerbate lead release (Dodrill and Edwards 1995). Large variations in the pH of the water of
the Washington D.C. service area have been identified as a contributing factor to the high lead
concentrations from 2000-2004 (U.S. Environmental Protection Agency 2007a).
Phosphate inhibitors can effectively mitigate lead concentrations in distribution systems.
The form in which the phosphate is added is vitally important. Orthophosphate and associated
forms (e.g., zinc orthophosphate) are the most effective inhibitors (Cantor et al. 2003; Churchill
et al. 2000; Edwards et al. 1999; Schock and Lytle 2010). Orthophosphate inhibitors can result
in the precipitation of insoluble lead phosphate solids that limit lead release to solution. The
dissolution rate of the lead carbonate hydrocerussite was significantly decreased by the addition
of orthophosphate in a study conducted over a pH range of 7.5-10 and various dissolved
inorganic carbon concentrations, and electron microscopy imaging confirmed the formation of a
low solubility lead(II) phosphate solid (Giammar et al. 2010). Orthophosphate was also found to
mitigate the release of dissolved lead from pipes during stagnant conditions (Xie and Giammar
2011). In contrast to orthophosphate, polyphosphates form dissolved complexes with lead and
can increase dissolved lead concentrations in distribution systems (Cantor 2006; Dodrill and
Edwards 1995; Edwards and McNeill 2002; McNeill and Edwards 2002; Schock et al. 2005a).
The benefits of orthophosphate as a corrosion inhibitor are most significant for systems
with relatively low alkalinity or pH, conditions at which lead concentrations are not controlled
by lead carbonate solids. In a survey of utilities, the most pronounced benefits of orthophosphate
addition were seen in waters with less than 30 mg/L alkalinity as CaCO3 (Dodrill and Edwards
1995). Recommended orthophosphate doses can range from less than 1 mg/L to as much as 4.5
mg/L as P (Edwards et al. 1999). The addition of orthophosphate to the treated water in
Washington D.C. has effectively lowered dissolved lead concentrations to below the action level
(U.S. Environmental Protection Agency 2007a; b).
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
CHAPTER 2
FACTORS CONTROLLING THE FORMATION OF LEAD(IV) OXIDES
OVERVIEW OF RESEARCH ON LEAD(IV) OXIDE FORMATION
The two polymorphs of lead(IV) oxide (PbO2), scrutinyite (α-PbO2) and plattnerite (βPbO2), have been widely observed as corrosion products in drinking water distribution systems
(Schock and Giani 2004; Schock et al. 2005b). The formation of lead(IV) oxide (PbO2) from
lead pipes is a two-step process. First, lead(0) metal is oxidized to lead(II) corrosion products,
like lead(II) oxides and carbonates. In the presence of free chlorine, these lead(II) compounds
are further oxidized to lead(IV) oxide. Figure 2.1 shows the possible formation pathways of
PbO2 from lead(II) species. Water chemistry parameters that include pH, dissolved inorganic
carbon (DIC), and free chlorine, as well as the identities of the precursors can affect the
formation of PbO2. The primary aims of the task presented in this chapter were to (1) investigate
the effect of water chemistry on the extent and identity of PbO2 formation and (2) determine the
pathways of PbO2 formation from various precursors.
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure 2.1 Potential pathways of PbO2 formation from initial Pb(II) phases
MATERIALS AND METHODS
Materials
Five different materials were studied as starting materials for PbO2 formation. Elemental
lead(0) powder was used to identify the formation pathways of PbO2 through the Pb(0)-Pb(II)Pb(IV) processes. To elucidate the formation pathways of PbO2 from Pb(II) phases, four Pb(II)
compounds were chosen as starting phases. Dissolved lead(II) chloride (PbCl2) was used to
study PbO2 formation from the aqueous phase. The lead(II) solids massicot (β-PbO), cerussite
11
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
12 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
(PbCO3), and hydrocerussite (Pb(OH)2(CO3)2) were used to identify the formation of PbO2 from
lead(II) corrosion products that are frequently observed in distribution systems.
Lead(0) powder (99.9% purity) was purchased from Alfa-Aesar. Reagent grade massicot
was purchased from Aldrich. A PbCl2 solution was prepared using a reagent grade PbCl2 salt
from Fisher Scientific. Cerussite and hydrocerussite were synthesized in our lab using methods
described before (Noel and Giammar 2007). Briefly, hydrocerussite was synthesized by
simultaneous addition of 0.1 M NaHCO3 and 0.15 M Pb(NO3)2 solutions to ultrapure water while
maintaining pH 9.0 + 0.5 by periodic addition of 1.0 M NaOH. To make cerussite, 0.1 M
NaHCO3 and 0.1 M Pb(NO3)2 were simultaneously added to ultrapure water while maintaining
pH 7.0 + 0.5. The identities of the solids were confirmed by X-ray diffraction (XRD) (Figure
2.2), and the morphology of the solids was characterized using SEM (Figure 2.3). The specific
surface areas of lead(0) powder, massicot, cerussite, and hydrocerussite were measured using
BET-N2 adsorption to be 0.05, 0.21, 1.09, and 4.77 m2/g, respectively. Reagent grade Pb(NO3)2,
NaHCO3, NaOH, NaOCl solution, and HNO3 were purchased (Fisher Scientific). Ultrapure
water (resistivity >18.2 MΩ-cm) was used to prepare solutions.
PbO2 Formation Experiments
The conditions for PbO2 formation were examined in batch experiments to determine the
factors that resulted in pure scrutinyite, pure plattnerite, neither phase, or mixtures of the two
phases. Factors evaluated were the form in which lead was added (i.e., the precursor), the pH,
Figure 2.2 X-ray diffraction patterns of the materials used in this study together with
reference patterns from the International Centre for Diffraction Data (ICDD) database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 2: Factors Controlling the Formation of Lead(IV) Oxides | 13
and the concentrations of free chlorine and DIC (Table 2.1). Experiments were performed at
room temperature (21 + 1 oC) using 500-mL polypropylene batch reactors. Lead precursors were
added to the reactors to achieve a total lead concentration of 0.5 mM. The desired DIC
concentration was provided by the addition of NaHCO3. Then the pH was adjusted to the target
value and maintained subsequently by the periodic addition of HNO3 and NaOH. Finally,
aliquots of a stock solution of sodium hypochlorite were added to provide the target free chlorine
concentrations. For experiments with 4 and 20 mg Cl2/L free chlorine, the free chlorine
concentration was maintained at the initial value by periodic addition of sodium hypochlorite
solution; while for experiments with 42 mg Cl2/L free chlorine, the free chlorine concentration
was allowed to decrease from its initial value. Two different approaches to providing the free
chlorine were used because these experiments were initiated by different researchers at different
times. For selected conditions both approaches were used, and there were no differences in the
results. The relatively high free chlorine concentrations were used in order to generate sufficient
amounts of products for characterization. To prevent the photodegradation of free chlorine, the
reactors were covered with aluminum foil. To minimize the uptake of CO2 from the atmosphere,
the solutions were stirred in sealed bottles. For experiments with no DIC, the experiments were
conducted in an argon-filled glovebox with the atmosphere in contact with a 0.5 M NaOH
solution to absorb any remaining CO2. After 1, 7, and 28 days of reaction solid samples were
collected and centrifuged prior to XRD and SEM analysis. Aqueous samples were taken
periodically, filtered with 0.22 μm nitrocellulose membranes, and analyzed for free chlorine.
a b c d
Figure 2.3. Electron micrographs showing the morphology of pure (a) cerussite, (b)
hydrocerussite, (c) massicot and (d) lead(0) powder
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
14 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Table 2.1
Experimental conditions studied for the formation of PbO2
Property
Selected Conditions
Dissolved lead: PbCl2 solution
Lead metal: Pb(0)(s)
Massicot: PbO(s)
Starting phase
Cerussite: PbCO3
Hydrocerussite: Pb3(CO3)2(OH)2
pH
7.5, 8.5, 10
DIC
0, 20 mg C/L
Free chlorine
0, 4, 20, 42 mg/L as Cl2
Reaction time
1 day, 7 days, 28 days
Analytical Methods
XRD was performed on a Rigaku Geigerflex D-MAX/A diffractometer using Cu-Ka
radiation. Electron microscopy was performed on a JEOL 7001LVF field emission scanning
electron microscope. BET-N2 adsorption was performed on a BET-Autosorb instrument
(Quantachrome Instruments). Free chlorine concentrations were determined by the standard
DPD colorimetric method (4500-Cl Chlorine G) with a spectrophotometer (Perkin Elmer
Lambda 2S) (Clesceri et al. 1999). Solution pH was measured with a glass pH electrode and pH
meter (Accumet).
RESULTS AND DISCUSSION
Overview
The conditions of experiments completed and information on the solids present after
reaction times are summarized in Table 2.2. PbO2 only formed in the presence of free chlorine.
Because the lowest free chlorine concentration studied was 4 mg Cl2/L, any threshold value
required for PbO2 formation is below this concentration. This observation is expected since
PbO2 is a strong oxidant that is only stable in a system with high oxidation reduction potential
(ORP). The pH and DIC affected the identity of the PbO2 produced by forming lead(II)
carbonate intermediate solid phases. The extent and identity of PbO2 formation also depended
on the starting phases.
Products of Reaction with Massicot
Effect of pH and DIC
The pH and DIC are important water chemistry parameters and determine the alkalinity
of the water. A typical range of DIC in drinking water is 0 – 50 mg C/L (Schock 1989) and 0
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 2: Factors Controlling the Formation of Lead(IV) Oxides | 15
Table 2.2
Summary of the resulting solid phases
pH
7.5
8.5
10
1d
7 d 28 d 1 d
7d
28 d
1d
7d
28 d
M
S/P
S/P
M
S/P
S/P
4
0
M
M/O M/S/P
M M/S/P M/S/P
20
PbO
M/H M/C C/M M/H M/H/L H/L/M M/H M/H/L H/L/M
(massicot)
0
(0.5 mM)
M/H M/C
P
M/H M/HS/P P/S
M/H M/H/S S/P
20
4
M/C/H P/S
P/S M/H S/P/M
S/P M/H S/P/M
S/P
20
N
N
S/P
Cl
S/P
P
4
PbCl2
0
(0.5 mM)
N
S/P
S/P
N
P
P
20
C/S
S/P
S/P
H/C S/P
S/P
20
Cerussite
0
(0.5 mM)
C/S
P/S
P/S
C/S S/P
S/P
42*
C/H/S
S/P
S/P
H/S
S
S
20
Hydrocerussite
0
(0.5 mM)
C/H/S S/P
S/P
S
S
H/S
42*
E
E/C
E/C
E
E/H
E/H
E
E/H
H/E
0
Pb powder
E
E/C E/C/O E
E/H
E/P/S
E
E
E/H/S
20
4
(0.5 mM)
E
E/P/S
P/S
20
S: Scrutinyite, P: Plattnerite, H: Hydrocerussite, C: Cerussite, E: Elemental lead, M: Massicot; L: litharge; Cl:
Pb4O3Cl2·H2O; O: PbO2, which was only identified by SEM.
The phase identified in bold was the predominant solid phase in the mixture
* indicates the initial free chlorine concentration; in these experiments the free chlorine was allowed to decline from
its initial value.
Precursor
DIC
Free Chlorine
(mg C/L) (mg/L as Cl2)
and 20 mg C/L was selected in the present study. In the absence of DIC, a mixture of scrutinyite
and plattnerite formed with 20 mg Cl2/L free chlorine (Figure 2.4). A peak at 2θ of 28o that is
characteristic of scrutinyite appeared in the XRD patterns after 28 days at both pH 7.5 and pH
10. The characteristic peaks of plattnerite at 2θ of 25o and 32o also appeared. No intermediate
solid was observed (Figure B1 in the Appendices). This result demonstrated that a mixture of
scrutinyite and plattnerite formed directly from massicot in the absence of DIC. The higher
intensity of the peaks at 2θ of 25o, 28o, and 32o at pH 10 than pH 7.5 indicated that PbO2
formation was faster at pH 10. The presence of DIC accelerated the formation of PbO2. The
peaks of massicot disappeared in the presence of DIC, but massicot still coexisted with PbO2 in
the absence of DIC after 28 days of reaction (Figure 2.4). The presence of DIC may accelerate
PbO2 formation by either forming solid intermediates or soluble Pb(II)-carbonate complexes that
have been demonstrated to enhance the dissolution of lead(II) solids (Giammar et al. 2008).
The pH affected the identity of the PbO2 formed in the presence but not the absence of
DIC. In the absence of DIC, mixtures of plattnerite and scrutinyite formed with similar
abundance after 28 days of reaction at both pH 7.5 and 10. In the presence of 20 mg C/L DIC
after 28 days of reaction with 20 mg Cl2/L free chlorine, mixtures of plattnerite and scrutinyite
formed; plattnerite was more dominant at pH 7.5, while scrutinyite was dominant at pH 10
(Figure 2.4). A similar trend was observed using 4 mg Cl2/L free chlorine (Figure B2 in the
Appendices). A previous study showed that plattnerite formed at pH 7.5 and scrutinyite formed
at pH 10 from PbCl2 solution in the presence of DIC (Lytle and Schock 2005), which is
consistent with the observation in the present study. The effect of DIC may be attributed to the
formation of lead(II) carbonate intermediate solids (Figures B3 - 4 in Appendices). SEM images
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
16 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure 2.4 X-ray diffraction patterns of solids following reaction of massicot with 20 mg
Cl2/L free chlorine in the absence and presence of 20 mg C/L DIC after 28 days together
with reference patterns from the ICDD database
a. b.
Figure 2.5 Electron micrographs of PbO2 products from reaction of PbO (a) after 1 day at
pH 7.5 with no DIC and 20 mg Cl2/L free chlorine and (b) after 1 day at pH 10 with 20
mg/L DIC and 20 mg Cl2/L free chlorine.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 2: Factors Controlling the Formation of Lead(IV) Oxides | 17
showed that PbO2 (small spherical particles) grew on the surface of massicot in the absence of
DIC; while in the presence of DIC, hydrocerussite intermediate solids (plate-like crystals) were
observed, and PbO2 formed on the edges of hydrocerussite (Figure 2.5). The formation of
lead(II) carbonate intermediate solids may affect the PbO2 formation pathways, thus affecting the
phases of the PbO2 formed.
Effect of free chlorine
Free chlorine concentration affected the extent of PbO2 formation and its impact was
strongly DIC-dependent (Figure 2.6). Higher free chlorine concentrations enhanced the
formation of PbO2 in the presence of DIC. After 7 days of reaction at pH 7.5 with 4 mg Cl2/L
free chlorine and 20 mg C/L DIC, a small XRD peak at 2θ of 32o appeared that indicated the
formation of PbO2. However, the dominant solids were still massicot and cerussite, which
formed from the reaction of massicot with DIC. Increasing the free chlorine concentration to 20
mg Cl2/L resulted in a complete transformation to scrutinyite and plattnerite after 7 days of
reaction. In the absence of DIC, an opposite trend was observed with less PbO2 formed at the
higher free chlorine concentration. Even after 28 days of reaction with 20 mg Cl2/L free
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure 2.6 X-ray diffraction patterns of solids following reaction of massicot at pH 7.5 with
and without 20 mg C/L dissolved inorganic carbon with different free chlorine
concentrations (in mg/L as Cl2) after 7 and 28 days together with reference patterns from
the ICDD database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
18 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
a. b.
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure 2.7 Electron micrographs of solids from massicot reaction at pH 7.5 with 20 mg
Cl2/L free chlorine, no DIC, after 28 days. Panel b is a higher magnification image of a
section from panel a.
chlorine, the dominant solids were still massicot; while scrutinyite and plattnerite became
dominant with 4 mg Cl2/L free chlorine.
This impact of DIC on the effect of free chlorine on PbO2 formation may be attributed to
different formation mechanisms of PbO2 from massicot in the absence and presence of DIC.
PbO2 may form through two parallel pathways (Figure 2.1): ① direct oxidation of PbO(s) to
PbO2(s) (reaction 3.1) and ② oxidation of dissolved lead that was released to solution from
PbO(s) (reactions 3.2 – 3.3):
PbO(S) + HOCl = PbO2(s) + H+ + ClPbO(s) + 2H+ =Pb2+ + H2O
Pb2+ + HOCl + H2O = PbO2(s) + Cl- + 3H+
(3.1)
(3.2)
(3.3)
In the absence of DIC, a high free chlorine concentration may enhance the formation of PbO2
from the direct oxidation of PbO, which can be seen by the formation of a possible PbO2 layer on
the surface of PbO at 20 mg Cl2/L free chlorine (Figure 2.7). This layer may have prevented
further chlorine attack of PbO and decreased the dissolution rate of PbO, thus inhibiting the
further formation of PbO2. The presence of DIC may accelerate Pb(II) release from massicot by
forming lead(II) carbonate complexes, thus increasing the rate of reaction 3.2, which would
allow reaction 3.3 to proceed faster than reaction 3.1. The increased dissolution of PbO at high
DIC may prevent the accumulation of PbO2 on the PbO surface, thus allowing continued PbO
dissolution and greater production of PbO2. In addition, the presence of DIC may help form
lead(II) carbonate intermediate solids, which may change the pathways of PbO2 formation,
resulting in a different trend than that observed in the absence of DIC.
Products of Reaction with Elemental Lead(0) Powder
PbO2 could form from lead(0) powder in the presence of free chlorine. Reacting with 20
mg Cl2/L free chlorine, PbO2 started to form after 7 days of reaction at pH 8.5 with 20 mg C/L
DIC, and it became dominant after 28 days of reaction (Figure 2.8). Small spherical particles
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 2: Factors Controlling the Formation of Lead(IV) Oxides | 19
Figure 2.8 X-ray diffraction patterns of solids following reaction of lead(0) powder at pH
8.5 with 20 mg C/L dissolved inorganic carbon and 20 mg Cl2/L free chlorine after different
reaction time together with reference patterns from the ICDD database
a. b.
Figure 2.9. Electron micrographs of PbO2 products from reaction of lead(0) powder at pH
8.5 with 20 mg C/L DIC and 20 mg Cl2/L free chlorine after (a) 7 days and (b) 28 days of
reaction.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
20 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
were observed after 7 days of reaction, indicating the formation of PbO2 (Figure 2.9). Plate-like
crystals were also found after 7 days of reaction, suggesting that hydrocerussite may be an
intermediate solid in the formation of PbO2 at this condition. After 28 days of reaction, the
solids all became small spherical particles, which is consistent with the XRD pattern that PbO2
was dominant after 28 days of reaction.
The free chlorine concentration affected the extent of PbO2 formation (Figure 2.10).
When no free chlorine was added to the solution, no PbO2 formed after 28 days of reaction.
Hydrocerussite formation was observed, but the dominant solid was still lead(0). With 4 mg
Cl2/L free chlorine, a mixture of scrutinyite and plattnerite formed after 28 days, and these solids
coexisted with the lead(0). When the free chlorine concentration was 20 mg Cl2/L, a mixture of
scrutinyite and plattnerite formed after 28 days, and the peaks of lead(0) almost disappeared.
Products of Reaction with Dissolved Lead Chloride
PbO2 formed from lead chloride at all the conditions tested in the absence of DIC. PbO2
formed after 28 days of reaction even when the free chlorine concentration was maintained at a
value as low as 4 mg Cl2/L, which indicated that the threshold value of free chlorine
Figure 2.10 X-ray diffraction patterns of solids following reaction of lead(0) powder at pH
8.5 with 20 mg C/L dissolved inorganic carbon with different free chlorine concentrations
(in mg/L as Cl2) after 28 days together with reference patterns from the ICDD database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 2: Factors Controlling the Formation of Lead(IV) Oxides | 21
concentration for PbO2 formation was less than 4 mg Cl2/L. The pH affected the phase of PbO2
formed in the absence of DIC. After 28 days of reaction, a mixture of scrutinyite and plattnerite
formed at pH 7.5 and pure plattnerite formed at pH 10 (Figure 2.11). This observation may be
due to the different formation pathways of PbO2 at different pH values. At pH 7.5, no
intermediate solid was observed during the formation of PbO2 (Figure 2.12a), which indicated
that the formation was through an aqueous phase oxidation pathway. This pathway includes two
steps: first dissolved Pb(II) is oxidized to dissolved Pb(IV), and then PbO2 nucleates
homogenously and precipitates, which is shown as pathway 2B-3A in Figure 2.1. At pH 10 with
4 mg/L free chlorine, an intermediate lead(II) oxide chloride was observed after 1 day of reaction
but disappeared after 7 days (Figure 2.12b). This intermediate solid may affect the overall PbO2
formation process, and the formation of pure plattnerite after 28 days of reaction may be due to
the specific trends in solid-solid interfacial free energies.
Products of Reaction with Lead(II) Carbonates
The lead(II) carbonates cerussite and hydrocerussite are often found as corrosion
products in pipe scales. They were also observed as intermediate solid phases in the formation
Figure 2.11 X-ray diffraction patterns of solids following reaction of lead chloride solution
in the absence of dissolved inorganic carbon with different free chlorine concentrations (in
mg/L as Cl2) after 28 days together with reference patterns from the ICDD database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
22 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure 2.12 X-ray diffraction patterns of solids following reaction of lead chloride solution
in the absence of dissolved inorganic carbon (a) at pH 7.5 with 20 mg Cl2/L free chlorine
and (b) at pH 10 with 4 mg Cl2/L free chlorine together with reference patterns from the
ICDD database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 2: Factors Controlling the Formation of Lead(IV) Oxides | 23
of PbO2 when DIC was present in the system (Lytle and Schock 2005). Cerussite and
hydrocerussite may transform into one another if the pH of the solution changes; the general
trend is that lower pH favors the formation of cerussite while higher pH leads to hydrocerussite.
Cerussite and hydrocerussite may affect the phases of PbO2 formed and their effect on the
formation of PbO2 was examined in the absence of DIC as a function of pH.
Reaction with Hydrocerussite
The pH affected the phase of PbO2 formed from hydrocerussite (Figure 2.13). At pH 10,
no intermediate solids were observed during the process of PbO2 formation, and pure scrutinyite
formed after 28 days of reaction with both 20 mg Cl2/L free chlorine and 42 mg Cl2/L free
chlorine. This result is consistent with previous studies conducted by Liu and co-workers that
found that scrutinyite formed from hydrocerussite (Liu et al. 2008). At pH 7.5 a phase
transformation from hydrocerussite to cerussite was observed after 1 day of reaction (Figure B5
in the Appendices), and at this pH PbO2 could then form from either cerussite or hydrocerussite,
which may explain why pure scrutinyite did not form. After 28 days, a mixture of scrutinyite
and plattnerite formed, and scrutinyite was more dominant than plattnerite.
Figure 2.13. X-ray diffraction patterns of solids following reaction of hydrocerussite in the
absence of dissolved inorganic carbon with different free chlorine concentrations (in mg/L
as Cl2) after 28 days together with reference patterns from the ICDD database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
24 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Reaction with Cerussite
A mixture of scrutinyite and plattnerite formed from cerussite after 28 days of reaction
(Figure 2.14). Except for the condition at pH 7.5 with 42 mg Cl2/L free chlorine in the absence
of DIC, scrutinyite was more dominant than plattnerite. A previous study observed that pure
plattnerite formed from cerussite at pH 6.7 – 7.8 (Lytle and Schock 2005). The formation of a
mixture of scrutinyite and plattnerite in the present study but not pure plattnerite may be due to
the partial transformation of cerussite to hydrocerussite. A phase transformation from cerussite
to hydrocerussite was observed at pH 10 after 1 day of reaction (Figure 2.15a), and PbO2 (small
roughly spherical particles) formed on the surface of hydrocerussite (hexagonal plate-like
crystals). At pH 7.5, although no hydrocerussite was observed with XRD, a hydrocerussite
shape-preserved cluster of PbO2 particles was observed after 7 days of reaction (Figure 2.15b),
indicating that at least some transformation from cerussite to hydrocerussite had occurred.
Therefore, at the conditions studied, PbO2 could have formed from both cerussite and
hydrocerussite, which makes it difficult to determine the formation products from pure cerussite.
Figure 2.14 X-ray diffraction patterns of solids following reaction of cerussite in the
absence of dissolved inorganic carbon with different free chlorine concentrations (in mg/L
as Cl2) after 28 days together with reference patterns from the ICDD database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 2: Factors Controlling the Formation of Lead(IV) Oxides | 25
a b Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure 2.15 Electron micrographs of the solid from cerussite reaction at (a) pH 10, 20 mg
Cl2/L free chlorine, and no DIC after 1 day, and (b) pH 7.5, 42 mg Cl2/L free chlorine, and
no DIC after 7 days
Summary of Lead(IV) Oxide Formation
The extent and identity of PbO2 formed were affected by water chemistry and the
precursors. PbO2 is a strong oxidant that only formed in the presence of free chlorine. Starting
from lead(0) powder, intermediate lead(II) solid phases were observed in the formation process
of PbO2, suggesting that the formation of PbO2 from Pb(0) is a two-step process that involves the
oxidation of Pb(0) to Pb(II) followed by the oxidation of Pb(II) to PbO2.
Starting from Pb(II) phases, the formation of PbO2 from different precursors at different
conditions is summarized in Figure 2.16. As shown in Figure 2.1, PbO2 could form from Pb(II)
phases through several pathways. If PbO2 formed from an aqueous phase oxidation pathway
(2B-3A in Figure 2.1), as illustrated in the experiments using dissolve lead chloride as a starting
phase at pH 7.5 with no DIC, mixtures of plattnerite and scrutinyite formed after 28 days of
reaction. If Pb(II) solids were involved in the process of PbO2 formation, then the mechanisms
of PbO2 formation can be more complicated. Pb(II) solids may be directly oxidized to PbO2
through a solid state oxidation pathway (2A in Figure 2.1) or serve as substrates in the
heterogeneous nucleation of PbO2 (1A-2B-3B in Figure 2.2). Generally if Pb(II) solids were
used as starting phases or were generated as intermediate solids, then mixtures of plattnerite and
scrutinyite formed. However, due to the specific trend in solid-solid interfacial free energies,
some Pb(II) solids did prefer to form a specific phase of PbO2. For example, lead(II) oxide
chloride favored plattnerite; while if hydrocerussite was used as the starting phase and no other
intermediate Pb(II) solids formed, then scrutinyite was the only product.
PbO2 solids have frequently been found on lead pipes in drinking water distribution
systems using free chlorine as a residual disinfectant. Our findings suggest that PbO2 formed in
the presence of free chlorine and that the formation rate of PbO2 increased with increasing free
chlorine concentrations with DIC. The threshold value above which the formation of PbO2 is
favored is less than 4 mg Cl2/L. In actual distribution systems the free chlorine concentration is
typically 0.5 to 1.5 mg Cl2/L, and the formation of PbO2 is still thermodynamically favorable and
has been observed on pipes. Although plattnerite is more thermodynamically stable than
scrutinyite, mixtures of plattnerite and scrutinyite formed at most conditions, indicating that the
systems do not reach equilibrium and that kinetics play an important role in PbO2 formation.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
26 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
pure scrutinyite
PbCl2 (aq)
pH 7.5; no DIC; Cl2 = 4, 20 mg/L
mixture
pH 10; no DIC
Cl2 = 4, 20mg/L
lead oxide chloride
pure plattnerite
pure scrutinyite
pH 10; DIC 20 mg C/L massicot
Cl2 = 4, 20 mg/L
hydrocerussite
massicot
pH 7.5, 10; no DIC; Cl2 = 4, 20 mg/L
pH 7.5; DIC 20 mg C/L
Cl2 = 20 mg/L
pH 7.5; DIC 20 mg C/L
Cl2 = 4 mg/L
massicot
cerussite
massicot
cerussite
pH 10; no DIC; Cl2 = 20, 42 mg/L
pH 7.5; no DIC
Cl2 = 20, 42 mg/L
mixture
pure plattnerite
pure scrutinyite
hydrocerussite
cerussite
mixture
hydrocerussite
pure plattnerite
pure scrutinyite
pH 10; no DIC hydrocerussite
Cl2 = 20, 42 mg/L cerussite
cerussite
mixture
pH 7.5; no DIC; Cl2 = 20, 42 mg/L
pure plattnerite
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure 2.16 Formation products from Pb(II) starting phases (dashed lines indicate
pathways that were not observed to occur in this study). When mixtures of scrutinyite
and plattnerite were produced, the vertical positions of the lines qualitatively indicate
their relative abundances. When intermediate solid phases were produced prior to
PbO2 formation, they are indicated.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
CHAPTER 3
DISSOLVED LEAD CONCENTRATIONS IN SOLUTIONS CONTAINING
LEAD(IV) OXIDE
OVERVIEW OF LEAD(IV) OXIDE SOLUBILITY
In the presence of free chlorine to promote the maintenance of lead in the +IV oxidation
state, PbO2 equilibrated in water should achieve stable dissolved lead concentrations.
Theoretically the stable dissolved lead concentrations should represent the equilibrium solubility
of PbO2, and if equilibrium can be achieved, then predictions of lead concentrations may be
made using equilibrium solubility calculations. However, it was not clear prior to this project
whether or not equilibrium could be achieved for aqueous PbO2 suspensions at near neutral pH.
At the high ORP provided by free chlorine, the dissolution of PbO2 is affected by two
types of reactions. The first is the non-reductive dissolution of PbO2, which releases Pb(IV) to
the solution, and can be expressed by Reactions 1 – 4 in Table 3.1. The total dissolved lead
Pb(IV) can be calculated as the sum of all the aqueous Pb(IV) species:
[Pb(IV)]diss =[Pb 4+ ]+[PbO32- ]+[PbO 4-4 ]
(3.1)
The second process is the reductive dissolution of PbO2, which may occur even in the presence
of free chlorine, that describes equilibrium between PbO2 and dissolved Pb(II) species (Reaction
5 in Table 3.1). In the system without DIC, the total dissolved Pb(II) can be expressed by:
[Pb(II)]diss = [Pb2+] + [PbOH+] + [Pb(OH)2] + [Pb(OH)3-] + [Pb(OH)42-]
Table 3.1
Reactions and rate constants of PbO2 dissolution
No
Reaction
logKeq
+
4+
1
-8.26
α- PbO2(s) + 4H → Pb + 2H2O
+
4+
2
-8.91
β- PbO2(s) + 4H → Pb + 2H2O
24+
+
3
-23.04
Pb + 3H2O → PbO3 + 6H
-63.8
4
Pb4+ + 4H2O → PbO44- + 8H+
2+
+
5
3.98
Pb + HOCl + H2O → PbO2(s) + Cl + 3H
2+
+
+
6
-7.71
Pb + H2O → PbOH + H
2+
+
7
-17.12
Pb + 2H2O → Pb(OH)2(aq) + 2H
2+
+
8
-28.06
Pb + 3H2O → Pb(OH)3 + 3H
22+
+
9
-39.70
Pb + 4H2O → Pb(OH)4 + 4H
20
2+
10
6.478
Pb + CO3 → PbCO3
222+
11
9.938
Pb + 2CO3 → Pb(CO3)2
+
2+
+
12
13.20
Pb + CO3 + H → PbHCO3
+
13
-7.60
HOCl → H + OCl
Benjamin = (Benjamin 2002)
MINEQL = (Schecher and Mcavoy 1998)
27
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
(3.2)
Source
Calculated
Calculated
Calculated
Calculated
Calculated
Benjamin
Benjamin
Benjamin
Benjamin
MINEQL
MINEQL
MINEQL
Benjamin
28 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Both pathways would contribute lead release to drinking water. Using the reactions and
equilibrium constants in Table 3.1 the calculated dissolved lead concentration as a function of
pH is shown in Figure 3.1. Above pH 6 Pb(IV) species become dominant and the dissolved lead
concentration increases with increasing pH. It should be noted that the predicted dissolved lead
concentrations in equilibrium with PbO2 are extremely low with concentrations always less than
10-9 M (0.20 µg/L) over the pH range of 4-11.
The low predicted concentrations are caused by the very low values for the equilibrium
constants for the dissolution of either of the two PbO2 solids (reactions 1-2). Caution should be
used when applying these values to the prediction of dissolved lead concentrations in solutions
with conditions relevant to drinking water distribution systems. The values reported in the table
were calculated from published values of Gibbs free energies of formation (G0f,i ) for the
relevant reactants and products (Pourbaix 1974; Risold et al. 1998); however, these Gibbs free
energies were determined from electrochemical potential measurements made at very acidic and
very basic conditions (Glasstone 1922) and not from any solubility experiments or experiments
conducted at near neutral pH. Further details of the calculation approach are presented in
Appendix A. The lack of any direct measurements of PbO2 solubility at conditions relevant to
drinking water distribution was a major motivation for Task 2.
Both of the PbO2 polymorphs, scrutinyite and plattnerite, have been observed in drinking
Figure 3.1 Predicted dissolved lead concentration in equilibrium with scrutinyite in the
presence of 2.8 * 10-5 M HOCl and Cl- based on reactions and equilibrium constants in
Table 3.1.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 3: Dissolved Lead Concentrations in Solutions Containing Lead(IV) Oxide | 29
water distribution systems. These two polymorphs have different calculated solubility constants,
so they may release different amounts of lead over long stagnation time. The primary objectives
of Task 2 were to (1) determine lead release from scrutinyite and plattnerite over long stagnation
times at different pH values and (2) gain knowledge of the equilibrium solubility of the two PbO2
polymorphs.
MATERIAL AND METHODS
Materials
Two PbO2 polymorphs were used. Scrutinyite was synthesized using the method
discussed below. Plattnerite was purchased (Acros) and its purity was confirmed by XRD.
Reagent grade PbCl2, NaHCO3, NaOH, NaOCl solution, and concentrated HNO3 were purchased
(Fisher Scientific). Ultrapure water (resistivity >18.2 MΩ-cm) was used to prepare solutions.
Synthesis of Scrutinyite
Scrutinyite was synthesized by oxidizing hydrocerussite with a sodium hypochlorite
(NaOCl) solution at pH 10 in the absence of DIC. Different hydrocerusite concentrations were
tested (0.5 to 5 mM), and preliminary results showed that pure scrutinyite only formed when
Figure 3.2 X-ray diffraction patterns of scrutinyite synthesized in this study together with
reference patterns from the ICDD database
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
30 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
starting with 0.5 mM hydrocerussite. Hydrocerussite was prepared using the method described
in Chapter 2. The synthesis experiments were conducted in 2-L polypropylene batch reactors at
room temperature (21+1 oC). Solid hydrocerussite was first added to the reactor to make a total
lead concentration of 0.5 mM. The pH was adjusted to 10.0 + 0.2 and then an NaOCl solution
was solution was added to the reactor to provide a free chlorine concentration of 40 mg/L as Cl2.
The free chlorine concentration was measured and maintained at its initial value by periodic
addition of the NaOCl stock solution. The key to forming pure scrutinyite was to maintain the
pH in the experiments. If the pH dropped below 8, then hydrocerussite would partially transform
to cerussite, which could lead to the formation of plattnerite impurities.
After 1 week of reaction, the solids were separated by centrifugation, washed with 1 mM
HNO3 and then with ultrapure water. The solids were then dialyzed for 48 hours using a
regenerated cellulose dialysis membrane, and the dialysis water was changed periodically with
fresh ultrapure water. The solids were freeze-dried and the identity of the solids was confirmed
using XRD (Figure 3.2). The morphology of the solids was determined using SEM (Figure 3.3).
Figure 3.3 Electron micrographs showing the morphology of scrutinyite
PbO2 Dissolution Experiments
Bench-scale experiments were conducted to determine the lead release from PbO2 over
long stagnation time and to acquire knowledge of the solubility of PbO2. A set of batch
experiments were conducted in 500-mL polypropylene reactors at room temperature (21+1 oC).
Factors evaluated included the form of the solid added (i.e. plattnerite vs. scrutinyite), pH, and
the presence of DIC (Table 3.2). For experiments in the presence of DIC, a 0.5 M NaHCO3
stock solution was used to adjust the DIC to 50 mg C/L. The free chlorine concentration was
then adjusted to above 2 mg Cl2/L and maintained subsequently by periodic addition of an
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 3: Dissolved Lead Concentrations in Solutions Containing Lead(IV) Oxide | 31
aliquot of a NaOCl stock over the course the 14-21 day experiments. The high free chlorine
concentration in the present study should be able to provide a sufficiently high ORP to prevent
the reductive dissolution of PbO2. The pH was adjusted to the target values and maintained by
the periodic addition of HNO3 or NaOH. Finally plattnerite or scrutinyite was added to the
reactors to make the solid concentration 50 mg/L. The reactors were covered with aluminum foil
to prevent the photodegradation of free chlorine. To get rid of possible side reactions, no buffers
were added to the system. To minimize the uptake of CO2 from the atmosphere, the solutions
were stirred in sealed bottles. For experiments with no DIC, the experiments were conducted in
an argon-filled glovebox with the atmosphere in contact with a 0.5 M NaOH solution to absorb
any remaining CO2.
Experiments were conducted for 14 to 21 days. Samples were taken periodically and
analyzed for dissolved lead and free chlorine concentrations. The samples for dissolved lead
analysis were first filtered through 0.22 µm polyethersulfone (PES) syringe filters, acidified to
2% HNO3, and preserved prior to analysis. Each experimental condition was run in duplicate.
Conditions/#
Table 3.2
Experimental conditions studied for the solubility of PbO2
1
2
Starting solid
plattnerite
Scrutinyite
50
50
6, 7.5, 8.5
7.5, 8.5
0, 50
0
Free Chlorine ( mg Cl2/L)
> 2 mg Cl2/L
> 2 mg Cl2/L
Experimental time (days)
21
14
Solid concentration (mg/L)
pH
DIC (mg C/L)
Analytical Methods
Dissolved lead (Pb) concentrations were measured by inductively coupled plasma mass
spectroscopy (ICP-MS) on an Agilent 7500ce instrument. The method detection limit for lead
was determined to be 0.05 µg/L. The approach followed that of Standard Method 3125. The
instrument was calibrated with standards (at least 7) with concentrations from 0.05 to 100 µg/L.
For samples with concentrations above the highest standard, dilutions were prepared. Free
chlorine concentrations were determined by the standard DPD colorimetric methods (4500-Cl G)
with a spectrophotometer (PerkinElmer Lambda XLS+) (Clesceri et al. 1999). Solution pH was
measured with a glass pH electrode and pH meter (Accumet).
RESULTS AND DISCUSSION
Effect of Total Lead Concentration on the Formation of Scrutinyite
To acquire enough pure scrutinyite for the solubility experiments, a set of experiments
were conducted to investigate the effect of total lead concentrations on the formation of
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
32 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
scrutinyite from hydrocerussite precursor at pH 10 in the absence of DIC. Results (Figure 3.4)
showed that the identity and purity of the PbO2 that formed strongly depended on the total lead
concentration. With 0.5 mM lead as hydrocerussite, 5 days of reaction led to the formation of
pure scrutinyite. When the total lead concentration increased to 1 mM, the characteristic peaks
of plattnerite at 25o and 32o in addition to those of scrutinyite were observed in the XRD pattern.
These peaks were also observed with total lead concentrations higher than 1 mM. Therefore, the
critical value for converting hydrocerussite to pure scrutinyite and not to mixtures of scrutinyite
and plattnerite is between 0.5 mM to 1 mM of lead at pH 10. From the discussion in Chapter 2, a
mixture of scrutinyite and plattnerite formed from an aqueous oxidation pathway. To avoid the
impurity of plattnerite and its potential effect on the lead release from scrutinyite, scrutinyite for
the solubility test was synthesized using 0.5 mM hydrocerussite as the starting phase.
Figure 3.4 X-ray diffraction patterns of solids following reaction of hydrocerussite at pH 10
without DIC and with a free chlorine:lead ratio of 1.2:1 together with reference patterns
from the ICDD database
Lead Release from Plattnerite
Effect of pH
Even in the presence of free chlorine, measurable amounts of lead may be released from
PbO2 after long stagnation times at conditions relevant to drinking water. Batch experiments
were conducted to determine lead release from plattnerite in the presence of free chlorine at pH
6.0 to 8.5. For experiments at pH 7.5 and 8.5, dissolved lead concentrations increased with
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 3: Dissolved Lead Concentrations in Solutions Containing Lead(IV) Oxide | 33
Figure 3.5 Dissolved lead concentrations with time in plattnerite batch dissolution
experiments. Experiments were performed at 50 mg/L plattnerite and 2 mg Cl2/L free
chlorine with (closed symbols) and without (open symbols) 50 mg C/L DIC at pH values of
(a) 6.0, (b) 7.5 and (c) 8.5. In all panels dashed lines represent the lead action level of 15
μg/L.”-A” and “-B” represent duplicate experiments.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
34 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
increasing time, and reached plateaus after 10 days of reaction (Figure 3.5). The total lead
concentrations only exceeded the lead action level (0.015 mg/L) after 10 days at pH 8.5 and 4
days at pH 7.5. At pH 6.0, lead concentrations were more variable and did not increase with
time. The lead concentrations were much lower at pH 6.0 than at pH 7.5 and 8.5. The
fluctuation of the lead concentrations may be partially due to the fluctuation of the system pH
since the system was poorly buffered due to the absence of DIC, and it was hard to maintain the
pH at the target value.
If the systems were at equilibrium, then the plateau lead concentrations may represent
equilibrium between PbO2 and the solution. The stable lead concentrations achieved in the batch
experiments were compared to the equilibrium lead concentrations calculated using the
published thermodynamic constants (Figure 3.6). The equilibrium calculation and the
measurements in this study do both predict a trend of increasing equilibrium lead concentrations
with increasing pH above pH 6. Clearly at all pH values studied, the measured equilibrium lead
concentrations were orders of magnitude higher than the predicted values. The large discrepancy
between the calculated and measured equilibrium solubility may be attributed to several reasons.
First the equilibrium constants used to calculate the solubility were derived from measurements
at very acidic and basic conditions, and those measurements were for electrochemical and not
solubility experiments. The thermodynamic constants for PbO2 and Pb(IV) species calculated
from published Gibbs from energies of formation may not be applicable to conditions relevant to
drinking water distribution.
Figure 3.6 Predicted (solid and dash lines) and measured (triangular points) equilibrium
solubility of PbO2 in the presence of 2 mg Cl2/L free chlorine and with no DIC.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 3: Dissolved Lead Concentrations in Solutions Containing Lead(IV) Oxide | 35
The second reason may be due to the presence of dissolved Pb(II) species. Although
PbO2 should be very stable in the presence of free chlorine and dissolved Pb(II) species should
be negligible, the kinetics of the reduction and re-oxidation reactions may be such that
appreciable reduction of PbO2 by water may still be occurring. The dissolved Pb(II)
concentration would then be controlled by the balance of the rates of the following two reactions:
PbO2(s) + 2H+ = Pb2+ + 0.5O2(aq) + H2O
HOCl + Pb2+ + H2O = PbO2 + Cl- + 3H+
(3.3)
(3.4)
If reaction 3.3 is faster than reaction 3.4, then aqueous Pb(II) species will be released to the
solutions even in the presence of free chlorine. The measured dissolved lead concentration will
then represent a balance between reactions 3.3 and 3.4, and the intrinsic equilibrium solubility of
PbO2 may not be able to be acquired. A previous study observed that 5 – 10 µg/L dissolved
Pb(II) was released from PbO2 even in the presence of free chlorine, which was in agreement
with this statement (Lin and Valentine 2009). The measured dissolved lead concentrations in the
present study that were orders of magnitude higher than the predicted values combined with the
dissolved Pb(II) detected in the previous study suggest that this second kinetically-based reason
may be the primary cause of the discrepancy between the measured and predicted lead
concentrations.
Effect of DIC
DIC is abundant in drinking water, and its effect on lead release from plattnerite over
long stagnation times was determined at pH 6.0 to 8.5 (Figure 3.5). Because carbonate is a good
complexing ligand for Pb(II), it may increase Pb(II) release from PbO2. However, in the present
study, DIC had very little effect on lead release at pH 6.0; while at pH 7.5 and 8.5, less lead was
actually released in the presence of 50 mg C/L DIC over the 21 days of reaction than in the
absence of DIC. The effect of DIC may be obscured by the effect of pH. When DIC was present
it served as a pH buffer and the system pH only fluctuated within 0.2 pH units of the target pH
values. While in the absence of DIC, the system was unbuffered and the pH fluctuated to a much
larger extent, around 0.5 pH units of the target pH values, which could have caused the lead
concentrations to be more variable. Overall, the results suggest that DIC has either no effect or a
minimal effect on long term lead release from plattnerite in the presence of free chlorine.
Lead Release from Scrutinyite
Lead release from scrutinyite (α-PbO2) was determined at pH 7.5 and 8.5 with no DIC in
the presence of free chlorine (Figure 3.7). Dissolved lead concentrations increased with time
until reaching plateau values after 10 days of reaction. The lead concentrations were above the
lead action level only after 6 days of reaction at both pH 7.5 and 8.5. Similar to the case of
plattnerite, the measured dissolved lead concentrations from scrutinyite were orders of
magnitude higher than those calculated from published thermodynamic constants (Figure 3.6),
suggesting that the lead release was controlled by the kinetics of reactions 3.3 and 3.4 rather than
by the intrinsic equilibrium solubility of scrutinyite. Since the measured dissolved lead
concentrations were several orders of magnitude higher than predicted using published
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
36 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
thermodynamic data for Pb(IV) species, it is reasonable to assume that nearly all of the measured
dissolved lead was present as aqueous Pb(II) species. After 14 days of reaction, more lead was
released at pH 8.5 than at pH 7.5.
Lead release from scrutinyite was compared to that from plattnerite (Figure 3.6). At
both pH 7.5 and 8.5, dissolved lead concentrations in systems equilibrated with scrutinyite were
similar to those in systems equilibrated with plattnerite. Scrutinyite and plattnerite are the two
polymorphs of PbO2, and they have similar thermodynamic constants. It was expected that the
equilibrium solubility of these two polymorphs would be similar. However, since the dissolved
lead concentrations may be controlled by the kinetics of reactions 3.3 and 3.4, these results
suggest that scrutinyite and plattnerite may also have similar dissolution kinetics.
Summary of Lead(IV) Oxide Solubility
The long term lead release from the two PbO2 polymorphs scrutinyite and plattnerite was
determined in batch reactors in the presence of free chlorine. Experiments were initially
designed to measure the equilibrium solubility of PbO2 with the assumption that PbO2 is stable in
solutions containing free chlorine. However, the measured dissolved lead concentrations were
orders of magnitude higher than the predicted values, suggesting that the lead release was not
controlled by the equilibrium solubility of PbO2 but rather by the kinetics of the reductive
dissolution of PbO2 by water and the re-oxidation of Pb(II) to PbO2. Despite being higher than
originally predicted, the stable lead concentrations in the presence of free chlorine were still
lower than the action level of 15 µg/L at pH 6 and only exceeded it after long stagnation times at
pH 7.5 and 8.5.
Although it is challenging to measure the equilibrium solubility of PbO2 directly from the
experiments, our results can provide an indication of the extent of lead release from PbO2 in
systems containing free chlorine with very long stagnation times. Most of these stagnation times
(4 days or longer) are much larger than those that would normally be encountered for a lead
service line. Consequently, for most lead service lines with PbO2 scales, the lead concentrations
will be determined by dissolution rates and not by equilibrium solubility. The long-term extent
of lead release increased with increasing pH, and plattnerite and scrutinyite released similar
amounts of lead. The presence of DIC had a limited effect on the extent of lead release from
plattnerite. Since dissolved lead concentrations in drinking water are controlled by the kinetics
of PbO2 dissolution rather than the equilibrium solubility of PbO2, it is crucial to determine the
dissolution rate of PbO2 at various water chemistry conditions and to elucidate the pathways and
mechanisms of PbO2 dissolution. The next chapter presents the results of a systematic study of
the dissolution rates of PbO2.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 3: Dissolved Lead Concentrations in Solutions Containing Lead(IV) Oxide | 37
Figure 3.7 Dissolved lead concentrations with time in scrutinyite batch dissolution
experiments. Experiments were performed at 50 mg/L scrutinyite, 2 mg Cl2/L free
chlorine, no DIC, and pH values of (a) 7.5 and (b) 8.5. In all panels dashed lines represent
the lead action level of 15 μg/L. ”-A” and “-B” represent duplicate experiments.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
CHAPTER 4
DISSOLUTION RATES OF LEAD(IV) OXIDE
OVERVIEW OF RESEARCH ON LEAD(IV) OXIDE DISSOLUTION RATES
Information about the rates and mechanisms of PbO2 dissolution are important for
estimating lead concentrations that may occur in water in contact with pipe scales that contain
this corrosion product. Because PbO2 is often found to coexist with lead(II) corrosion products
in drinking water distribution systems (Schock and Giani 2004), it is clear that these systems are
not at equilibrium. Consequently, information about rates and not equilibrium solubility is
needed. The dissolution rate of PbO2 is affected by water chemistry parameters including pH
and DIC. PbO2 is such a strong oxidant that it is not stable in pure water. Previous studies
indicated that the presence of reductants in drinking water could increase the rate and extent of
PbO2 dissolution. These reductants include natural organic matter (NOM), Fe2+, Mn2+, I-, and
Br-, (Dryer and Korshin 2007; Lin et al. 2008; Lin and Valentine 2008b; 2010; Shi and Stone
2009a; b). The majority of these studies were not designed to provide quantitative information
on dissolution rates, which led to a knowledge gap that the present project sought to fill.
The reductive dissolution of PbO2 can be described by a mechanism involving coupled
chemical reduction and detachment steps. The combined process of reduction and detachment is
referred to here as dissolution. In the first step the reductants attach to the surface of PbO2 and
reduce the surface Pb(IV) species to Pb(II) species; in the next step the surface Pb(II) species
detach and are released to the solution. Water chemistry parameters can influence the rate of
dissolution of PbO2 by affecting these two steps. Carbonate from DIC can act as a complexing
ligand for Pb(II) that may accelerate the detachment step of the overall dissolution process. As
noted above, reductants that are naturally present in the water may increase the rate of reduction
of Pb(IV) to Pb(II). Free chlorine is a strong oxidant that may either inhibit the reduction of
surface Pb(IV) species to Pb(II) species or re-oxidize the Pb(II) species that are released to water.
Orthophosphate is often added as a corrosion inhibitor in water treatment, and it may inhibit the
dissolution of PbO2 by either adsorbing and blocking surface sites of PbO2 or forming a low
solubility lead(II) phosphate solid. The primary aims of the task whose results are presented in
this chapter were to (1) determine the dissolution rates of PbO2 as a function of water chemistry,
(2) elucidate the mechanisms of the reductive dissolution of PbO2 at conditions relevant to
drinking water, and (3) quantify the inhibitory effects of free chlorine and orthophosphate on the
dissolution rate of PbO2.
MATERIALS AND METHODS
Materials
PbO2 was purchased (Acros), and XRD patterns confirmed that the PbO2 was pure
plattnerite (Figure 4.1). The PbO2 solids had primary sizes of around 100 nm (Figure 4.2), and
the specific surface area of the PbO2 was 3.6 m2/g as measured by BET-N2 adsorption. Reagent
grade KH2PO4, KI, NaNO3, NaHCO3, NaOH, MOPS, NaOCl solution, NH4Cl, and concentrated
HNO3 were purchased (Fisher Scientific). Ultrapure water (resistivity >18.2 MΩ-cm) was used
to prepare solutions.
39
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
40 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Figure 4.1 X-ray diffraction patterns of the PbO2 used in this study together with the
reference pattern from the ICDD database
Figure 4.2 Electron micrographs showing the morphology of pure plattnerite
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 41
PbO2 Dissolution Experiments
Summary of Experimental Conditions
PbO2 dissolution was examined as a function of water chemistry. The parameters
investigated were pH, DIC, the presence of iodide, free chlorine, and orthophosphate. Four sets
of experiments were conducted for different purposes in understanding the overall mechanisms
of PbO2 dissolution and the effect of potential inhibitors (Table 4.1). In Set A the dissolution
rates of PbO2 in water were determined in the presence of DIC to provide a baseline. In Set B
the dissolution of PbO2 was investigated in the presence of iodide as a chemical reductant and
carbonate as a complexing ligand to probe the pathways and mechanisms of its reductive
dissolution. This set integrated the effects of both the reductant and a complexing ligand on the
dissolution rate. In Set C the dissolution of PbO2 was investigated as a function of free chlorine
concentration to determine its inhibitory effect and to probe for a critical concentration that
might be able to inhibit dissolution. In Set D orthophosphate was added to the solution to
evaluate its effect on lead release from PbO2.
Flow-through Experiment Set-up
Continuously stirred tank reactors (CSTRs) were used to quantify the dissolution rates of
PbO2 at room temperature (21 ± 1 ºC) (Figure 4.3). The volume of each reactor was 84 mL, and
PbO2 was loaded to the reactor to a concentration of 1 g/L. A 0.22 μm mixed cellulose filter
membrane was used to seal the reactor and prevent the loss of the solid from the reactor. The
influent flow was pumped into the reactor using a peristaltic pump (Cole-Parmer) with a flow
rate of 2.8 mL/min, so the hydraulic residence time of the reactor was 30 minutes. The effluent
pH was monitored and aqueous samples were periodically collected over 24 hours and preserved
for dissolved lead analysis. Selected samples were further filtered using 0.02-μm PES syringe
filters and their close agreement with the 0.22-µm filtered samples suggested that the filtrate
from the 0.22-µm filters represents dissolved lead and did not include any colloidal lead.
The influents were prepared in 10-L plastic (Tedlar) bags to minimize the exchange of
CO2 between the atmosphere and the solution. Ultrapure water was purged of CO2 by sparging
with N2 and then pumped into the bags. NaHCO3 was added to provide the desired
concentrations of DIC. For the experiments without DIC at pH 7.6, an aliquot of 0.5 M 3-(Nmorpholino) propanesulfonic acid (MOPS) solution was injected to achieve a concentration of 1
mM. MOPS was selected as a pH buffer due to its low affinity for metal complexation and
relatively little effect on PbO2 dissolution. A 1.0 M NaNO3 solution was then injected to the
bags to set the ionic strength at 0.01 M. The pH was adjusted to the target values by addition of
concentrated HNO3 or freshly prepared 0.5 M NaOH solutions. For Set B iodide was used as a
reductant and its concentration was adjusted by addition of an aliquot of 10 mM KI stock
solution. For Set C free chlorine was provided by a NaOCl stock solution. For Set D instead of
free chlorine, an aliquot of a 200 mg Cl2/L chloramine stock solution was added to provide a
target chloramine concentration of 2 mg Cl2/L. The stock chloramine solutions were prepared by
mixing volumes of 6% (w/w) NaOCl and 2500 mg NH3/L NH4Cl solutions in ultrapure water.
This mixture provides a 0.79 Cl2:N molar ratio that can simulate the conditions in drinking water
distribution systems and premise plumbing. Under these conditions the dominant form of
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
42 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Set
A
Table 4.1
Experimental conditions studied for the dissolution rate of PbO2
Influent Solution
Conditions
pH: 5.7, 6.7, 7.6, 8.5
Ultrapure water
DIC: 50 mg C/L
B
Reductant (iodide) and complexing ligand
solution (DIC)
pH: 5.7, 6.7, 7.6, 8.5
DIC: 0, 10, 50, 200 mg C/L
Iodide: 0, 1, 2, 5, 10, 20, 100 µM
C
Free chlorine
pH: 5.7, 6.7, 7.6, 8.5
DIC: 50 mg C/L
Free chlorine: 0.2, 1 mg Cl2/L
Orthophosphate
pH: 7.6, 8.5
DIC: 0, 50 mg C/L
Orthophosphate: 0, 1 mg P/L
Chloramine: 2 mg Cl2/L
D
chloramine is monochloramine (NH2Cl). Then the orthophosphate concentration was adjusted to
1 mg P/L by adding a volume of 10 mM KH2PO4 stock solution. Solutions with free chlorine or
chloramine were shielded from light by aluminum foil to minimize their decomposition.
Information on Influent Composition
Four sets of experiments were conducted to evaluate the effect of different parameters on
the dissolution rate of PbO2 and to elucidate the pathways and mechanisms of PbO2 dissolution.
The effect of different parameters was examined by varying the influent composition. Four sets
Figure 4.3 Flow-through reactor for measuring dissolution rates
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 43
of experiments with different influents were selected (Table 4.2).
A: Ultrapure water: PbO2 is a strong oxidant that it is even not stable in water. The
reductive dissolution of PbO2 by water to release Pb(II) is thermodynamically favorable.
Experiments were conducted at pH 5.7 to 8.5 in the presence of 50 mg C/L DIC, a concentration
that is in the range observed for drinking water. A total of 4 conditions were examined.
Table 4.2
Conditions studied for PbO2 dissolution
pH
Sets A & C
Free chlorine
(mg Cl2/L)
5.7
6.7
7.6
8.5
0
1
2
3
4
0.2
23
26
28
1
22
25
27
24
Iodide concentration (µM)
Set B
1
2
10
5.7
DIC (mg C/L)
50
5
6.7
DIC (mg C/L)
50
6
13
0
pH
5
7.6
8.5
DIC (mg C/L)
DIC (mg C/L)
100
11
12
14
10
50
20
15
7
8
9
10
200
16
10
20
50
17
18
200
19
21
pH
Set D*
DIC (mg
C/L)
0
7.6
8.5
Orthophosphate ( mg P/L)
Orthophosphate (mg P/L)
0
1
0
1
29
33
30
34
50
31
35
32
36
Numbers are experiment identification numbers for each condition. Gray boxes are conditions
that were not examined.
*Experiments in Set D were all conducted in the presence of 2 mg Cl2/L chloramine.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
44 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
B: Solution with complexing ligand and reductant: The reductive dissolution of PbO2
can be described by a coupled reduction-detachment pathways. The presence of a chemical
reductant or a complexing ligand may enhance the overall dissolution of PbO2. Iodide was
selected as a model reductant in the present work with a concentration range of 1 – 100 µM.
Iodide is present in natural waters and is capable of reducing PbO2 at drinking water conditions
to form aqueous iodine species (Fuge and Johnson 1986). Although iodide may not be the most
significant reductant in most drinking waters, the observation of its impacts on PbO2 dissolution
can provide insights into the overall reductive dissolution of PbO2, which may be induced by
other reductants in drinking water that include natural organic matter and Fe(II) and Mn(II). DIC
is abundant in drinking water with typical concentrations of 0 – 50 mg C/L. Carbonate is a good
complexing ligand for Pb2+ that can increase the dissolution rate of PbO2 by forming lead(II)carbonate complexes. Carbonate was selected as a model complexing ligand with the DIC
concentration of 0 – 200 mg C/L. Experiments were conducted at pH 5.7 to 8.5. A total of 17
conditions were examined.
C: Solution with free chlorine: Free chlorine is a strong oxidant and can provide a high
ORP that may prevent the reductive dissolution of PbO2. Two free chlorine concentrations were
evaluated for their potential inhibitory effect on the dissolution rate of PbO2. Experiments were
conducted at pH 5.7 – 8.5 in the presence of 50 mg C/L. A total of 7 conditions were examined.
D: Solution with orthophosphate: Switching the disinfectant from free chlorine to
chloramine can decrease the ORP of the system and cause the reductive dissolution of PbO2.
Adding orthophosphate is considered as a strategy to mitigate lead release and has been adopted
by several water utilities. In this set of experiments the dissolution rate of PbO2 was determined
in the presence and absence of 1 mg P/L orthophosphate to evaluate the inhibitory effect of
orthophosphate and to determine the mechanisms of the inhibition. To simulate the conditions
typically found in drinking water with orthophosphate addition, experiments were performed
with 2 mg Cl2/L chloramine in the presence of 50 mg C/L DIC at pH 7.6 and 8.5. A total of 8
conditions were examined.
Dissolution Rate Determination Method
The dissolution rates were determined for the CSTRs by operating the reactors until they
reached steady-state behavior. As compared to batch experiments, the flow-through approach
avoids the accumulation of reaction products and minimizes the effect of any initial labile phases
on the dissolution rates (Samson et al. 2000). Performing a mass balance on the dissolved lead
in the reactor, the following relationship can be acquired:
dC
VR . =Q.Cin -Q.Cout +rexp .VR
(4.1)
dt
where:
rexp = measured net dissolution rate of PbO2 (mol/L·min)
VR = volume of the reactor (mL)
Q = flow rate (mL/min)
Cin, Cout = influent and effluent lead concentration (M)
The systems approach steady state (i.e. dC/dt = 0) after 10 to 20 residence times of
operation (Xie et al. 2010). The influent is lead free (i.e. Cin = 0), and the effluent concentrations
at steady state are denoted as Css, which was calculated as the average concentration from at least
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 45
8 consecutive samples that did not vary by more than 20% and that spanned at least 8 residence
times. Therefore the experimental rate can be determined as:
Q  Css Css
rexp =
=
(4.2a)
VR
t res
where:
tres = the hydraulic residence time (i.e. VR/Q) (min)
And the rate can be normalized to surface area:
Q  Css
Css
R exp =
=
(4.2b)
VR  A  [solids] t res  A  [solids]
where:
Rexp = measured net dissolution rate of PbO2 normalized to surface area (mol/m2·min)
A = specific surface area of PbO2 (m2/g)
[solids] = concentration of PbO2 solids in the reactor (g/L)
Pb(II) Adsorption Experiments
The adsorption of aqueous Pb(II) ions onto the surface of PbO2 was determined at room
temperature (21 + 1 oC) using 150-mL polypropylene batch reactors. A 10 mg Pb/L lead(II)
chloride stock solution was added to achieve a total lead concentration of 200 µg/L (0.96 µM).
Then a 0.5 M NaHCO3 stock solution was added to provide a 50 mg C/L of DIC. The pH was
adjusted to 4.5 to 8.5 with HNO3 or NaOH. Finally, PbO2 solids were added to provide a solid
concentration of 0.1 g/L. The system was well-mixed. After 2 hours, aqueous samples were
collected, filtered with 0.22 µm PES syringe filters, acidified to 2% HNO3, and preserved for
dissolved lead analysis. The difference between the total lead added to the reactors and the
dissolved lead after 2 hours represents the amount of Pb(II) that had adsorbed to the PbO2.
Analytical Methods
Dissolved lead (Pb) concentrations were measured by inductively coupled plasma mass
spectrometry (ICP-MS) on an Agilent 7500ce instrument. Solids remaining at the end of
selected experiments were collected, centrifuged, and characterized by XRD and SEM. A
Rigaku Geigerflex D-MAX/A diffractometer with Cu-Ka radiation was used for XRD
measurements. Electron microscopy was performed on a JEOL 7001LVF field emission
scanning electron microscope. BET-N2 adsorption was performed on a BET-Autosorb
instrument (Quantachrome Instruments). Free chlorine, combined chlorine, and orthophosphate
concentrations were determined by the standard DPD and ascorbic acid colorimetric methods
(4500-Cl G and 4500-P E) with a spectrophotometer (PerkinElmer Lambda XLS+) (Clesceri et
al. 1999). Solution pH was measured with a glass pH electrode and pH meter (Accumet).
RESULTS AND DISCUSSION
Overview
Four sets of flow-through experiments were conducted to determine the effect of different
parameters on the dissolution rate of PbO2. All experimental conditions studied and the results of
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
46 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
the steady-state dissolved lead concentrations and the dissolution rates are listed in Table 4.3. In
DIC-buffered ultrapure water, the dissolution rate of PbO2 increased with decreasing pH values.
The presence of free chlorine inhibited the dissolution of PbO2, and the dissolution rate
decreased with increasing free chlorine concentrations. The presence of orthophosphate
inhibited the dissolution of PbO2 in the presence of DIC and chloramine, which may be due to
the adsorption of orthophosphate to the surface of PbO2 to block the active sites for PbO2
dissolution.
The presence of the chemical reductant iodide significantly enhanced the dissolution of
PbO2, and the dissolution rate increased with increasing iodide concentrations and decreasing
pH. The addition of a small amount of DIC (10 mg C/L) increased the dissolution rate of PbO2,
and further increases in the DIC concentration from 10 to 200 mg C/L had little effect on the
dissolution rate of PbO2. These observations suggested that chemical reduction was usually the
rate-controlling step in the dissolution of PbO2.
PbO2 Dissolution at Baseline Conditions – Ultrapure Water (Set A)
The dissolution rate of PbO2 in water with 50 mg C/L DIC was determined at pH 5.7 to
8.5. The effluent dissolved lead concentrations used to determine the dissolution rate and the pH
are shown as a function of residence times in Figure 4.4. The effluent lead concentrations were
more variable during the first 10 – 20 residence times, which may be due to fast initial
dissolution of labile phases and more reactive sites. The systems reached steady state by 40
residence times, and the effluent lead concentrations were used to calculate the dissolution rate
of PbO2. Compared to the initial dissolution rate, the steady-state dissolution rate of PbO2 may
be more representative in describing the dissolution of PbO2 in actual distribution systems since
the steady-state dissolution avoids the accumulation of any reaction products and minimizes the
effect of the presence of labile phases.
The dissolution rate of PbO2 increased with decreasing pH values (Figure 4.5). The
dissolution rate increased only slightly from pH 8.5 to 6.7, while it increased four times in going
from pH 6.7 to 5.7. Results suggested that for a typical drinking water pH range (7.6 and 8.5 in
the present study), some Pb(II) would be released to drinking water from the dissolution of PbO2,
but the dissolution rate is expected to be low. However, if the system is acidic, then the
dissolution rate of PbO2 would increase significantly. Although pH 5.7 is outside of the range
anticipated for drinking water distribution systems, it might be present in dead zones of
distribution networks or regions of low flow where biofilms grow or nitrification occurs (Zhang
et al. 2009). Investigation at this pH was also helpful in establishing the overall trend with pH.
The effect of pH may be caused by several reasons. First, the reductive dissolution of
PbO2 by water is thermodynamically favorable and pH may affect the electrochemical driving
force for PbO2 reduction. The reductive dissolution of PbO2 by water can be described by
Reaction 4.3:
PbO2(s) + 2H+ = Pb2+ + 0.5O2(aq) + H2O
(4.3)
Decreasing the pH would increase the electrochemical driving force (ΔEH) for Reaction 4.3.
Based on the Nernst equation and published EH° values for the relevant half reactions (Reactions
1 and 2 in Table 4.4), ΔEH will increase by 0.059V for a 1 pH unit decrease. Increasing the ΔEH
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 47
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
48 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 49
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
50 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 51
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
52 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Figure 4.4 Effluent lead concentrations (■ and ▲) and pH (□ and ) from CSTRs as a
function of the number of hydraulic residence times (tres = 30 min) in the 50 mg C/L DICbuffered water at pH (a) 5.7, (b) 6.7, (c) 7.6, and (d) 8.5. Duplicate experiments
(represented by rectangles and triangles) were conducted for each condition.
would increase the electrochemical driving force for PbO2 reduction, thus increasing the
dissolution rate of PbO2.
Table 4.4
Thermodynamics of PbO2 dissolution
Reaction
EH0 (V)
No
PbO2 reduction:
PbO2 + 4H+ + 2e- = Pb2+ + 2H2O
1.45
1
Water oxidation:
2H2O = O2(aq) + 4H+ + 4e-
-1.27
2
-0.62
3
1.48
4
-
-
Iodide oxidation:
2I = I2(aq) + 2e
Free chlorine reduction:
HOCl + H+ + 2e- = H2O + Cl-
EH0 was calculated using the thermodynamic data from (Stumm and Morgan 1996).
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 53
Figure 4.5 The dissolution rate of plattnerite in the 50 mg C/L DIC-buffered water at pH
5.7 to 8.5 using flow-through reactors with 1 g/L plattnerite and a residence time of 30 min.
Error bars represent one standard deviation from duplicate experiments.
Figure 4.6 Adsorption of 0.96 µM Pb(II) onto 0.1 g/L PbO2(s) in the presence of 50 mg C/L
dissolved inorganic carbon (DIC) as a function of pH.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
54 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
A second reason may be related to the adsorption of Pb2+ ions onto the surface of PbO2.
Pb could adsorb to the PbO2 surface at neutral or slightly basic environments (Figure 4.6),
which may passivate the PbO2 surface by blocking the active sites for the dissolution of PbO2.
At acidic conditions Pb2+ does not adsorb as completely to the PbO2 surface such that more
surface sites would be unoccupied and available for reductive dissolution of PbO2.
2+
Effect of Chemical Reductant and Complexing Ligand on PbO2 Dissolution (Set B)
Effect of Iodide and pH
The reductive dissolution of PbO2 by iodide can be described by the half reactions
involving PbO2 reduction and iodide oxidation (Reactions 1 and 3 in Table 4.4). The overall
redox reaction between PbO2 and iodide can be expressed as:
PbO2 + 2I- + 4H+ = Pb2+ + I2 + 2H2O
(4.4)
The presence of 10 µM iodide increased the dissolution rate of PbO2 at all pH values
studied (Figure 4.7). The presence of a reductant can significantly affect the stability of PbO2.
The dissolution rate increased with decreasing pH, and the apparent reaction order with respect
Figure 4.7 Effect of pH on the dissolution rate of plattnerite with 50 mg C/L DIC in the
absence and presence of 10 μM iodide using flow-through reactors with 1 g/L plattnerite
and a residence time of 30 min. Error bars represent one standard deviation from
duplicate experiments.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 55
to H+ concentrations was 0.61 over the range of pH 5.7 to 8.5 in the present study. This
observation was comparable to that of a previous study that measured the dissolution rate of
PbO2 by iodide and found that the reaction order with respect to H+ was 0.66 for pH 6.0 – 8.0
(Lin et al. 2008).
The effect of pH may be caused by its effects on the electrochemical driving force for
PbO2 reduction by iodide and by its impact on iodide adsorption onto the PbO2 surface.
Increasing the difference in the electrochemical potentials for PbO2 reduction and iodide
oxidation (ΔEH) would increase the electrochemical driving force for PbO2 reduction, and
increase the dissolution rate. Based on the Nernst equation and published EH° values for the
relevant half reactions (Reactions 1 and 3 in Table 4.4), the ΔEH of Reaction 4.4 will increase by
0.118V for a 1 pH unit decrease. In addition, more iodide would be expected to be adsorbed
onto the surface of PbO2 at lower pH since this behavior is typically observed for anion
adsorption. The greater extent of PbO2 surface coverage by iodide at lower pH may result in
faster dissolution of PbO2 at lower pH.
Effect of Iodide Concentration
The effect of different iodide concentrations on the dissolution rate of PbO2 was
evaluated at pH 7.6 and 8.5 in the presence of 50 mg C/L DIC. Increasing iodide concentrations
generally increased the dissolution rate of PbO2. A threshold value above which the dissolution
Figure 4.8 Effect of iodide concentration on the dissolution rate of plattnerite with 50 mg
C/L DIC at pH 7.6 and 8.5 using flow-through reactors with 1 g/L plattnerite and a
residence time of 30 min. Error bars represent one standard deviation from duplicate
experiments.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
56 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
of PbO2 was accelerated was observed for experiments both at pH 7.6 and 8.5 (Figure 4.8). At
pH 7.6, the PbO2 dissolution rate with 1 µM iodide was comparable to that in the absence of
DIC; increasing the iodide concentrations to 2 µM doubled the dissolution rate, and the
dissolution rate further increased with increasing iodide concentrations. The threshold iodide
concentration that enhanced the dissolution of PbO2 at pH 7.6 was between 1 and 2 µM. At pH
8.5, the presence of 5 µM iodide did not increase the dissolution rate of PbO2. PbO2 started to
dissolve faster with 10 µM iodide, indicating that the threshold value was between 5 and 10 µM
at pH 8.5. The higher threshold value at pH 8.5 than at pH 7.6 was qualitatively consistent with
equilibrium calculations based on Reaction 4.4 that suggest that higher iodide concentrations are
required at higher pH to result in comparable PbO2 dissolution extents as at lower pH.
Trends of the steady-state PbO2 dissolution rate may be different than those acquired
from the initial dissolution of PbO2 in batch experiments. Due to the presence of labile phases
and highly reactive sites, the initial dissolution of PbO2 may be more variable and much faster
than the steady-state dissolution determined over longer time periods. Using the flow-through
reactors, the dissolution products were continuously flushed out of the reactor. This approach
allowed us to measure the long term steady-state dissolution of PbO2, which was more
representative of real distribution systems and was not affected by any labile phases. Indeed, the
effluent lead concentrations over the first several residence times were more variable than the
steady-state concentrations. For iodide concentrations above 5 µM the initial dissolution rates
were also higher than the ultimate stable steady-state dissolution rate achieved, which suggests
the dissolution of some labile phases during the initial period of the CSTR experiments (Figure
Figure 4.9 X-ray diffraction patterns of the plattnerite after 24 hours of reaction at pH 7.6
with 50 mg C/L DIC and different iodide concentrations together with reference patterns
from the ICDD database. Some cerussite formation is apparent when the iodide
concentration was 100 μM.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 57
C3 in the Appendices).
Faster initial dissolution of PbO2 with high iodide concentrations may release sufficient
Pb(II) to form Pb(II)-precipitates. For example, when starting with 100 µM iodide at pH 7.6, the
initial lead release was fast and the system became supersaturated with respect to the Pb(II)carbonate cerussite (PbCO3). The XRD patterns for the solids after the experiment show the
formation of cerussite (Figure 4.9). Cerussite formation was also confirmed from the SEM
observation. Long-bar shaped crystals were only observed at the condition with 100 µM iodide,
suggesting the formation of cerrusite at this condition but not for lower iodide concentrations
(Figure 4.10). The formation of the secondary lead(II) precipitates can provide a sink for Pb(II)
species, and they may also change the available surface area of PbO2. Therefore for this highest
iodide concentration studied, the steady-state effluent lead concentrations do not represent the
dissolution rate of PbO2, but rather they indicate the balance between PbO2 dissolution and
cerussite formation. Consequently for the iodide concentrations above 20 µM, the dissolution
rates cannot be acquired from Equation 4.2.
a b c Cerussite
Figure 4.10 Electron micrographs of solids: (a) before reaction, and after 24 hours of
reaction in the CSTR with 50 mg C/L DIC and (b) 10 μM iodide at pH 7.6, and (c) 100 μM
iodide at pH 7.6.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
58 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Effect of Dissolved Inorganic Carbon
DIC is abundant in drinking water and carbonate is a good complexing ligand for Pb2+.
For example, at pH 7.6 with 50 mg C/L DIC, 93% of the dissolved lead is composed of Pb(II)carbonate complexes that include PbCO3(aq), Pb(CO3)2-, and PbHCO3+. The effect of different
DIC concentrations on the dissolution rate of PbO2 was determined with 5 or 10 μM iodide at pH
7.6 and 8.5. Experiments in the absence of DIC were conducted at pH 7.6 with 5 and 10 μM
iodide using 1 mM MOPS as a buffer. The inclusion of the buffer was critical to being able to
provide a stable pH for comparison with the carbonated system when no DIC was available to
buffer the pH. Organic buffers can accelerate the dissolution of PbO2 due to their chemical
reduction of PbO2. When compared with other organic buffers (MES or HEPES), MOPS was a
mild reductant and only slightly enhanced the dissolution of PbO2 (Figures C1 and C2 in the
Appendices). Even after accounting for the enhancement of MOPS on the dissolution of PbO2,
the presence of 50 mg C/L DIC still significantly accelerated PbO2 dissolution (Figure 4.11).
With either 5 or 10 μM iodide, the dissolution rates of PbO2 with 50 mg C/L DIC were nearly
twice as high as those without DIC. Carbonate from DIC can serve as a complexing ligand and
react with Pb2+ ions to form Pb(II)-carbonate complexes to enhance the dissolution of PbO2.
Figure 4.11 Dissolution rate of plattnerite at pH 7.6 in the absence and presence of 50 mg
C/L DIC with different iodide concentrations using flow-through reactors with 1 g/L
plattnerite and a residence time of 30 min. Error bars represent one standard deviation
from duplicate experiments.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 59
Although the presence of DIC increased the dissolution rate of PbO2, varying the DIC
concentrations from 10 to 200 mg C/L had little effect on the dissolution rate of PbO2 at both pH
7.6 and 8.5 with 10 μM iodide (Figure 4.12). The reductive dissolution of PbO2 can be described
as a coupled process of chemical reduction and detachment, and the presence of carbonate may
affect the detachment of Pb(II) from PbO2 surface. Our results suggest that once some DIC was
present in the system, the detachment step would not be rate-limiting in determining the overall
dissolution of PbO2 by iodide and that chemical reduction may play the most important role in
controlling the dissolution rates.
Comparison of Iodide with Different Reductants
The PbO2 dissolution rates acquired in the present study using flow-through reactors were
compared to those determined in batch reactors with other reductants found in drinking water
(Table 4.5). The presence of chemical reductants was proved to enhance the dissolution of PbO2
in all of the studies, but to different extents. In the present study, the dissolution rate of PbO2
was 4.7·10-9 mol·m-2·min-1 at pH 7.6 in the presence of 10 mg C/L DIC and 10 μM iodide, which
was comparable to the rate determined in a phosphate-buffered system in a previous study. Lin
and Valentine studied the initial dissolution of PbO2 by iodide using batch experiments. Based
on the general rate expression they provided, the dissolution rate of PbO2 at pH 7.0 with 10 μM
Figure 4.12 Effect of DIC on the dissolution rate of plattnerite with 10 μM iodide at pH 7.6
and 8.5 using flow-through reactors with 1 g/L plattnerite and a residence time of 30 min.
Error bars represent one standard deviation from duplicate experiments.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
60 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
iodide was 2.1·10-9 mol·m-2·min-1(Lin et al. 2008). NOM is widely present in drinking water
and was found to be able to reduce PbO2 (Dryer and Korshin 2007; Lin and Valentine 2008b). A
previous study investigated the effect of NOM on the dissolution rate of PbO2 at pH 7.0 with 20
mg/L NOM and acquired a dissolution rate of 4.0·10-9 mol·m-2·min-1 (Lin and Valentine 2008b),
which was comparable to the rate determined from iodide in the present work. If metal cations,
like Fe2+ and Mn2+ were present in drinking water, then they may accelerate the dissolution of
PbO2 more significantly than do iodide or NOM (Shi and Stone 2009a).
Generally, if the concentrations of the reductants discussed earlier are similar, then the
order of PbO2 dissolution rates can be summarized as NOM < I- < Mn2+, Fe2+. A more relevant
case is to consider the dissolution rate with the relative abundance of the different reductants. In
actual distribution systems, the Mn2+ concentrations may be as high as 0.2 mg/L (4 µM) (Cerrato
et al. 2006), the NOM concentration could be several mg C/L (Volk et al. 2002), and the typical
iodide concentration is below 20 μg/L (0.16 µM) (National Academy of Sciences 1980), which is
below the threshold value that could accelerate the dissolution of PbO2. Therefore iodide may
play a less important role than NOM or Mn2+ and Fe2+ in enhancing lead release from PbO2
dissolution in typical distribution systems. Nevertheless, iodide was useful as a model reductant
for elucidating the general mechanisms and pathways of the reductive dissolution of PbO2.
Table 4.5
Summary of the dissolution rate of PbO2 by different reductants.
Experimental Condition
Dissolution
PbO2
Reductant
Mode Rate (mol·mRef
DIC
Reductant
pH
conc.
2
·min-1)a
(M)
conc.
(mg/L)
(Dryer and
10 mg/L
-4
-9
NOM
7.0
10
36
Batch
4.8·10
Korshin
DOC
2007)
(Lin and
NOM
7.0
10-3
20 mg/Lb
4
Batch
4.0·10-9
Valentine
2008b)
(Shi and
7.8
0
40 μM
4.8
Batch
1.3·10-6
Stone
Fe2+
2009a)
(Shi and
Mn2+
7.8
0
20 μM
4.8
Batch
3.2·10-6
Stone
2009a)
7.0
0
500 μM
8.0
Batch
1.5·10-6
(Lin et al.
II7.0
0
10 μM
103
Batch
2.1·10-9c
2008)
Present
I7.6 8.3*10-4
10 μM
103
Flow
4.7·10-9
study
a
The dissolution rate was calculated from the data included in the figures of the cited
reference.
b
The carbon content of the NOM was 40%.
c
The dissolution rate was calculated using the rate expression provided in the reference (Lin
et al. 2008) applied to experimental conditions similar to those of the present study.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 61
Pathways of the Reductive Dissolution of PbO2
The reductive dissolution of a metal oxide, like Fe(III) or Mn(III, IV), can be described
by several steps, including the adsorption of the reductant to the surface of the metal oxide,
electron transfer between the metal oxide and the reductant, and detachment of the reduced metal
ions from the metal oxide surface (Hering and Stumm 1990; Stone and Morgan 1984). The
reductive dissolution of PbO2 will follow these general steps, and the following reactions can be
used to describe the reactions of PbO2 with iodide (Figure 4.13):
(a) adsorption of iodide to the surface of PbO2 to form a precursor surface complex:
k1

  Pb( IV ) I  H 2O
(4.5)
 Pb( IV )OH  I -  H  
k 1
(b) two one-electron transfers between the surface Pb(IV) species and iodide:
k2
fast
 Pb( IV ) I 
  Pb( III ) I 
  Pb( II )OH
(4.6)
(c) detachment of the Pb(II) species from the PbO2 surface to solution:
kd
 Pb ( II )OH 
  Pb ( IV )OH  Pb 2 
(4.7)
Steps a and b are affected by the iodide concentrations and step c would be accelerated by
the presence of a Pb(II)-complexing ligand. In the presence of DIC, changing the DIC
concentrations from 10 to 200 mg C/L had little effect on the dissolution rate of PbO2,
suggesting that step c was fast. Increasing the iodide concentrations accelerated the dissolution
of PbO2, indicating that steps a and b were rate-limiting. Since iodide is involved in both steps a
and b, they will be collectively referred to as “steps related to chemical reduction.” If the
chemical reduction steps are rate-limiting, then at a given pH the reductive dissolution of PbO2
follows Langmuir-Hinshelwood kinetics, and the dissolution rate can be expressed by:
Figure 4.13 Pathways of the reductive dissolution of PbO2 by iodide
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
62 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
R k'
k[ I  ]
1  k[I  ]
(4.8)
where k (μM-1) and k’ (mol·m-2·min-1) are rate constants at a given pH. When the iodide
concentration is low, the dissolution rate of PbO2 is first order with respect to iodide
concentrations. As the iodide concentration increases, the dissolution rate would reach a plateau.
This type of behavior was observed at pH 7.6 (Figure 4.8).
Employing least squares optimization of the experimental data to Equation 4.8, at pH 7.6
the optimal value of k’ was 99·10-10 mol·m-2·min-1 and of k was 8.1·10-2 μM-1; at pH 8.5 the
optimal values of k’ and k were 96·10-10 mol·m-2·min-1 and 1.0·10-2 μM-1. The model can
generally fit the data quite well as shown in Figure 4.14 which shows the comparison of the
model dissolution rate with the experimental dissolution rate. It is anticipated that the
mechanism and model proposed in the present work may also be applicable to interpreting the
reactions of PbO2 with other reductants in distribution systems.
Effect of Free Chlorine on PbO2 Dissolution (Set C)
Free chlorine is often used as a disinfectant in drinking water distribution systems. PbO2
is thermodynamically stable in systems containing free chlorine since free chlorine provides a
high ORP. The effect of free chlorine on the dissolution rate of PbO2 was examined over a pH
Figure 4.14 Model for PbO2 dissolution rate in the presence of iodide at pH 7.6 and 8.5. The
1:1 line is shown for reference.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 63
Figure 4.15 The dissolution rate of plattnerite with 50 mg C/L DIC in the absence and
presence of 1 mg Cl2/L free chlorine at pH 5.7 to 8.5 using flow-through reactors with 1
g/L plattnerite and a residence time of 30 min. Error bars represent one standard
deviation from duplicate experiments.
range from 5.7 to 8.5 with 50 mg C/L and 1 mg Cl2/L free chlorine. The presence of 1 mg Cl2/L
free chlorine significantly inhibited the dissolution of PbO2 (Figure 4.15). For pH 6.7 to 8.5, the
presence of free chlorine decreased the dissolution rate of PbO2 by more than an order of
magnitude.
Since the presence of 1 mg Cl2/L free chlorine effectively inhibited the dissolution of
PbO2, the possibility of a critical free chlorine concentration that can prevent the reductive
dissolution of PbO2 was investigated. A lower free chlorine concentration of 0.2 mg Cl2/L was
selected, and the dissolution rate of PbO2 was determined at pH 7.6 and 8.5 with 50 mg C/L DIC.
Although PbO2 dissolved a little bit faster with 0.2 mg Cl2/L free chlorine than with 1 mg Cl2/L
free chlorine, the dissolution rates with 0.2 mg Cl2/L free chlorine were still an order of
magnitude lower than those without free chlorine (Figure 4.16). These results demonstrated that
even the presence of 0.2 mg Cl2/L free chlorine dramatically inhibited the dissolution of PbO2
and maintained low dissolved lead concentrations. The threshold value of free chlorine
concentration that inhibited the dissolution of PbO2 was below 0.2 mg Cl2/L.
The inhibitory effect of free chlorine on the dissolution rate of PbO2 may be due to
several reasons. Free chlorine may provide a high enough ORP so that the reductive dissolution
of PbO2 was not thermodynamically favorable. But it is more likely that free chlorine affected
the kinetics of a rate-limiting step in the overall PbO2 dissolution process. Free chlorine may reoxidize the dissolved Pb(II) species released from PbO2 (Reactions 4.9 and 4.10), thus mitigating
the lead release to water.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
64 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
HOCl + Pb2+ + H2O = PbO2 + Cl- + 3H+
OCl- + Pb2+ + H2O = PbO2 + Cl- + 2H+
(4.9)
(4.10)
Effect of Orthophosphate on PbO2 Dissolution (Set D)
The effect of orthophosphate on the dissolution rate of PbO2 was investigated at pH 7.6
and 8.5 in the absence and presence of 50 mg C/L DIC with 2 mg Cl2/L chloramine. The
addition of 1 mg P/L orthophosphate inhibited the dissolution of PbO2 at pH 7.6, but the
orthophosphate had little effect on the dissolution rate of PbO2 at pH 8.5 (Figure 4.17). For the
experiments at pH 8.5 in the absence of DIC, the dissolution rate of PbO2 was actually higher in
the presence of orthophosphate. This observation may be due to the pH difference between the
conditions with and without orthophosphate. No buffer was used at pH 8.5 in the absence of
DIC and therefore the pH fluctuated more widely and was not always maintained at the target pH
value. The average pH was 8.69 to 8.85 in the absence of orthophosphate, and it was 8.07 to
8.44 in the presence of orthophosphate. Consequently, the dissolution rate acquired at this
condition may actually be caused by the pH but not by the orthophosphate.
Figure 4.16 Effect of free chlorine concentration on the dissolution rate of plattnerite with
50 mg C/L DIC at pH 7.6 and 8.5 using flow-through reactors with 1 g/L plattnerite and a
residence time of 30 min. Error bars represent one standard deviation from duplicate
experiments.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 65
Figure 4.17 Effect of orthophosphate concentration on the dissolution rate of plattnerite
with 0 and 50 mg C/L DIC at pH 7.6 and 8.5 using flow-through reactors with 1 g/L
plattnerite and a residence time of 30 min. The legend shows the orthophosphate
concentrations in mg P/L. Error bars represent one standard deviation from duplicate
experiments.
Figure 4.18. Conceptual model of the potential pathways of PbO2 dissolution in the
presence of orthophosphate.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
66 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Orthophosphate may inhibit the dissolution of PbO2 through several mechanisms (Figure
4.18). The lead(II)-phosphate solid hydroxylpyromorphite (Pb5(PO4)3OH) has a very low
solubility. When the dissolved lead released from PbO2 dissolution exceeds the equilibrium
solubility of hydroxylpyromorphite, hydroxylpyromorphite may precipitate and control the lead
concentration. Precipitation of hydroxylpyromorphite on the PbO2 surface may also block
reactive sites on the surface of PbO2 and prevent its reductive dissolution.
Hydroxylpyromorphite precipitates were observed at pH 7.6 with 1 mg P/L orthophosphate in
the absence of DIC (Figure 4.19c). The dissolution rate at this condition was lower than that in
the absence of orthophosphate, suggesting that precipitation of hydroxylpyromorphite inhibited
the dissolution of PbO2.
(a)
a (b)
b (c)
(d)
c d Figure 4.19. Electron micrographs of (a) plattnerite before reaction, (b) pure
hydroxylpyromorphite, (c) solids after reaction of plattnerite at pH 7.5, 0 mg C/L DIC, 1
mg P/L phosphate, and 2 mg Cl2/L monochloramine, and (d) solids after reaction of
plattnerite at pH 8.5, 50 mg C/L DIC, 1 mg P/L orthophosphate, and 2 mg Cl2/L
A second possible mechanism of orthophosphate inhibition of PbO2 dissolution involves
the adsorption of phosphate to the surface of PbO2. Adsorbed phosphate may block the reactive
sites for PbO2 reduction or detachment, thus passivating the PbO2 surface and mitigating lead
release. In the presence of DIC, the predicted equilibrium solubility of hydroxylpyromorphite is
higher than that without DIC due to the formation of soluble Pb(II)-carbonate complexes. At the
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Chapter 4: Dissolution Rates of Lead(IV) Oxide | 67
conditions with 1 mg P/L and 50 mg C/L DIC at pH 7.6 and 8.5, the steady-state effluent lead
concentrations did not reach the equilibrium solubility of hydroxylpyromorphite (Figure 4.20).
SEM observations also indicated that no secondary lead(II) solid phases formed following the
reaction of PbO2 with orthophosphate in the presence of DIC (Figure 4.19d), which is consistent
with the lead concentrations being undersaturated with respect to hydroxylpyromorphite.
Therefore the inhibitory effect of orthophosphate is probably due to adsorption and not
precipitation when DIC is present. Orthophosphate adsorption should follow the general trend of
anion adsorption with less extensive adsorption with increasing pH. Consequently, a larger
extent of the PbO2 surface would be covered by orthophosphate at pH 7.6 than at pH 8.5, which
is in agreement with the observation that orthophosphate had a larger inhibitory effect on the
dissolution of PbO2 at pH 7.6 than at pH 8.5.
Overall, there are two mechanisms through which orthophosphate can inhibit the release
of lead to solution during the reductive dissolution of PbO2. In the absence of DIC the formation
of the lead(II) phosphate hydroxylpyromorphite limited lead release to solution. With DIC
present, the inhibition of lead release to solution was probably caused by the adsorption of
orthophosphate to the PbO2 surface to block sites of reduction or dissolution.
Figure 4.20 Steady-state concentrations for plattnerite dissolution in the presence of 1 mg
P/L orthophosphate and predicted lead concentration in equilibrium with
hydroxylpyromorphite.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
68 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Summary of PbO2 Dissolution Rates
The dissolution rate of PbO2 was strongly dependent on water chemistry. For the
conditions studied, the pH was an important factor with dissolution rates always increasing with
decreasing pH. The presence of free chlorine or orthophosphate inhibited the dissolution of
PbO2, and free chlorine was more effective than orthophosphate at maintaining low dissolved
lead concentrations in the reactor effluents. Free chlorine concentrations as low as 0.2 mg Cl2/L
decreased the dissolution rate of PbO2 by more than an order of magnitude at conditions relevant
to drinking water; the presence of 1 mg P/L orthophosphate decreased the dissolution rate to a
lesser degree. The presence of iodide (i.e. a chemical reductant) and DIC (i.e. a complexing
ligand) accelerated lead release from PbO2, and iodide was more important in controlling the
overall dissolution because of the significance of chemical reduction as the rate-controlling step
in the overall PbO2 dissolution process for most conditions..
The reductive dissolution of PbO2 can be described by several sequential steps: (1) the
adsorption of the reductant to the surface of PbO2; (2) electron transfer between the surface
Pb(IV) species and the reductant, resulting in the reduction of the surface Pb(IV) to surface
Pb(II); and (3) detachment of the surface Pb(II) from the PbO2 surface. The effects of the
different water chemistry factors on PbO2 dissolution rates can be interpreted with respect to
their different impacts on the pathways of PbO2 dissolution. At a given pH, the presence of
iodide can accelerate steps (1) and (2), and the presence of DIC can accelerate step (3).
Collectively the iodide and DIC result in an enhancement of PbO2 dissolution. The presence of
free chlorine can inhibit the dissolution of PbO2 by re-oxidizing the Pb(II) ions released from
step (3) to form PbO2 solids. The presence of orthophosphate may inhibit the PbO2 dissolution
by either forming an insoluble Pb(II)-phosphate precipitate from the Pb(II) ions released from
step (3) or adsorbing to the surface of PbO2 to block reactive sites, which may then prevent the
surface from the attack of the chemical reductants or complexing ligands in steps (1), (2), or (3).
Varying the pH may affect the extent of the iodide and orthophosphate adsorption to the PbO2
surface, thus affecting the overall dissolution rate of PbO2.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
CHAPTER 5
SUMMARY AND CONCLUSIONS
SUMMARY OF PROJECT
The project involved three integrated tasks that focused on the formation and dissolution
of the important lead corrosion product PbO2. PbO2 formation was systematically examined in
Task 1. Batch experiments were performed to determine the extent of formation and identity of
PbO2 (plattnerite versus scrutinyite) as a function of important water chemistry parameters. The
results of this task were presented in Chapter 2. The chapter included the results of
characterization of the solid phases and interpretation of the PbO2 formation pathways. In Task
2, PbO2 dissolution was examined over long equilibration times in batch reactors. Lead release
from the two PbO2 polymorphs, scrutinyite and plattnerite, was determined in the presence of
free chlorine. The results of this task were presented in Chapter 3. In Task 3, continuously
stirred tank reactors were used to investigate the dissolution rate of PbO2 as a function of the
water chemistry parameters pH, DIC, iodide, free chlorine, and orthophosphate. The results of
this task were presented in Chapter 4. These results included dissolved phase analysis,
characterization of the solid phases, and description of the dissolution pathways and mechanisms
using a reaction-based framework.
FORMATION OF LEAD(IV) OXIDE
The extent of formation and identity of the PbO2 formed strongly depended on the water
chemistry and the lead-containing precursor. PbO2 only formed in the presence of free chlorine.
The threshold free chlorine concentration above which PbO2 could form was below 4 mg Cl2/L.
While PbO2 formed in the presence and absence of DIC, the presence of DIC accelerated PbO2
formation. The pH and DIC affected the PbO2 formation pathways by forming different
intermediate lead(II) solid phases, thus affecting the extent and phase of the PbO2 formed.
Starting from elemental lead(0) powder, intermediate lead(II) solid phases were observed,
confirming that the formation of PbO2 from Pb(0) is a two-step process that involves the
oxidation of Pb(0) to Pb(II) followed by the oxidation of Pb(II) to PbO2. Starting from the
lead(II) precursors, PbO2 formed through several pathways, including an aqueous phase
oxidation pathway and solid phase mediated pathways. Mixtures of plattnerite and scrutinyite
formed from the aqueous phase oxidation pathway. When Pb(II) solids were involved in the
process of PbO2 formation, mixtures of plattnerite and scrutinyite formed at most experimental
conditions. However, due to the specific trend in solid-solid interfacial free energies, some
Pb(II) solids result in the preferential formation of a specific phase of PbO2. Lead(II) oxide
chloride favored plattnerite. When no intermediate solids formed, hydrocerussite led to the
formation of scrutinyite.
SOLUBILITY OF LEAD(IV) OXIDE
The solubility of PbO2 was investigated in batch reactors that maintained a free chlorine
residual to provide a high ORP. Over long stagnation times, dissolved lead concentrations were
orders of magnitude higher than the values calculated from equilibrium constants derived from
69
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
70 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
published thermodynamic data. It is very likely that the dissolved lead measured was aqueous
Pb(II) rather than aqueous Pb(IV), and therefore the solubility of PbO2 equilibrated with only
aqueous Pb(IV) species cannot be determined experimentally. The results suggested that even in
the presence of free chlorine, the dissolved lead release from PbO2 still represents a balance of
two kinetic processes: reductive dissolution of PbO2 by water and re-oxidation of aqueous Pb(II)
species by free chlorine.
Although the measured concentrations did not match the extremely low predicted
concentrations, the final stable dissolved lead concentrations were still very low when free
chlorine concentrations were maintained; dissolved lead concentrations never exceeded the
action level at pH 6.0 even after several weeks of equilibration, and at pH 7.5 and 8.5 the lead
concentrations only exceeded the action level after four days or longer. Increasing the pH
increased the dissolved lead concentrations over long stagnation times. Relative to pH, DIC had
less impact on lead release from PbO2 in the presence of free chlorine. The two PbO2
polymorphs, scrutinyite and plattnerite, released similar amounts of dissolved lead over long
stagnation times.
DISSOLUTION RATE OF LEAD(IV) OXIDE
The dissolution rate of PbO2 was influenced by all five water chemistry parameters
tested. The pH consistently played a role in affecting dissolution rates with higher rates almost
always observed with decreasing pH. The pH may affect the distribution of the PbO2 surface
species and the pathways of the PbO2 dissolution.
Free chlorine or orthophosphate decreased the dissolution rate of PbO2, and free chlorine
was more effective than orthophosphate at inhibiting PbO2 dissolution. The presence of free
chlorine concentrations even as low as 0.2 mg Cl2/L decreased the dissolution rate by more than
an order of magnitude. Free chlorine may inhibit the dissolution of PbO2 by re-oxidizing the
Pb(II) species back to PbO2. The presence of 1 mg P/L orthophosphate slightly inhibited the
dissolution of PbO2 at pH 7.5. There are two mechanisms through which orthophosphate may
inhibit the release of lead to solution during the reductive dissolution of PbO2. In the absence of
DIC the formation of the lead(II) phosphate hydroxylpyromorphite limited lead release to
solution. With DIC present, the inhibition of lead release to solution was probably caused by the
adsorption of orthophosphate to the PbO2 surface to block sites of reduction or dissolution.
The presence of iodide and DIC accelerated the dissolution of PbO2. The reductive
dissolution of PbO2 can be described by several sequential steps: (1) the adsorption of the
reductant to the surface of PbO2; (2) electron transfer between the surface Pb(IV) species and the
reductant, resulting in the reduction of the surface Pb(IV) to surface Pb(II); and (3) detachment
of Pb(II) from the PbO2 surface. Iodide accelerated the steps related to the chemical reduction,
and increasing the iodide concentration increased the dissolution rate of PbO2. The presence of
DIC accelerated the detachment of Pb(II) from PbO2 surface by forming soluble Pb(II)-carbonate
complexes, thus enhancing PbO2 dissolution, but further increases in DIC from 10 to 200 mg
C/L had little impact on the dissolution rate of PbO2 in the systems containing iodide. Chemical
reduction was more important than detachment in controlling the overall dissolution of PbO2.
Although iodide may not be a significant reductant in many distribution systems, the overall
mechanism of reductive dissolution of PbO2 determined with iodide may be applicable to PbO2
dissolution as influenced by natural organic matter and other reductants.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
CHAPTER 6
RECOMMENDATIONS TO UTILITIES
DETERMINING WHETHER OR NOT LEAD(IV) OXIDES ARE PRESENT
Lead(IV) oxides (PbO2) can be an important component of corrosion products on pipe
scales for utilities that have lead service lines in their distribution systems and that currently use
or have used free chlorine as the secondary disinfectant. PbO2 can only form in systems with
free chlorine present; however, because of the low solubility of this phase and the complexity of
pipe scales, PbO2 may persist well after a switch from free chlorine to chloramine. The rate of
PbO2 formation and consequently the likely extent of PbO2 formation on lead service lines as
well are strongly affected by the water chemistry of the distribution system. Utilities should
assume that PbO2 is present on lead service lines if free chlorine is used; however, it is possible
that not all systems using free chlorine will have PbO2 present despite its predicted formation.
When samples of lead service lines become available (e.g., during replacements as part of water
main repair), characterization of the composition of the pipe scales can be performed to confirm
the presence of PbO2. Such identification is best accomplished using X-ray diffraction. The
susceptibility of PbO2 to reductive dissolution and its likely enrichment at the surface of the pipe
scale can make it a major contributor to lead release to water even when it is only present as a
minor constituent of a pipe scale.
The presence of PbO2 is a more important finding then the exact identity of the form of
PbO2 (plattnerite versus scrutinyite) that is present. Both plattnerite and scrutinyite can form on
lead pipes in the presence of free chlorine, and the specific solid that is formed is affected by
both the water chemistry and the precursor lead(II) phase from which it forms. The coexistence
of scrutinyite and plattnerite and previous observations of their co-occurrence with Pb(II) solids
indicates that the systems have not reached equilibrium and that rates play an important role in
PbO2 formation. Despite predictions that plattnerite should have a lower solubility than
scrutinyite, similar dissolved lead concentrations were achieved in aqueous solutions equilibrated
with the two solids.
MAINTAINING LOW DISSOLVED LEAD CONCENTRATIONS IN DISTRIBUTION
SYSTEMS THAT CONTAIN LEAD(IV) OXIDES
Lead(IV) oxides are only stable in solutions with a high oxidation-reduction potential
maintained by a free chlorine residual. In the absence of free chlorine, PbO2 dissolves through a
reductive dissolution process. For systems without free chlorine or for which free chlorine has
been depleted, dissolved lead concentrations will be controlled by the rate of the dissolution
reaction and not by equilibrium solubility. Even in the presence of free chlorine, while stable
low dissolved lead concentrations can be achieved, it is likely that those stable concentrations are
controlled by a balance of reductive dissolution and re-oxidation reactions and not by true
equilibrium. As a general rule equilibrium solubility calculations are not useful for predicting
dissolved lead concentrations in waters in contact with PbO2. Because rates play such an
important role in affecting lead concentrations, the flow regime (e.g., flowing water versus
stagnation) must also be considered when estimating lead concentrations in water that is in
contact with PbO2 as a corrosion product.
71
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
72 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
The dissolution rate of PbO2 is a very strong function of the water chemistry, and orders
of magnitude differences can occur in the rates depending on the composition. Even in the
absence of free chlorine, water compositions with moderate to high pH, low DIC, and the
absence of any chemical reductants (e.g., natural organic matter) can have such slow dissolution
of PbO2 that low lead concentrations are maintained over reasonable stagnation times. Similarly
for such systems, PbO2 could persist as a component of pipe scales for long periods of time
(years or longer) even after a switch from free chlorine to another disinfectant. Consequently,
utilities should still be cautious in altering water compositions in ways that could affect PbO2
stability even after secondary disinfection is switched from chlorination.
Utilities should be aware of conditions that could accelerate the dissolution of PbO2 and
the release of lead from this potential reservoir of unstable lead in scales on lead service lines.
This is particularly important for water distribution systems that have recently switched from
using free chlorine to chloramine as the secondary disinfectant and for regions of distribution
systems with high water ages such that the residual chlorine may have been completely
consumed. Dissolution rates of PbO2 will increase in response to process changes that lower the
pH of the water in the distribution system. The most significant parameter affecting PbO2
dissolution rates is the concentration and identity of species that can act as chemical reductants to
accelerate PbO2 dissolution. These species include natural organic matter, dissolved iron(II) and
manganese(II), and iodide. Measurements of these properties of the finished water can provide
information regarding the potential for lead release from PbO2 present on lead service lines.
Carbonate can also act to accelerate PbO2 dissolution, but since this effect is only eliminated at
very low DIC concentrations that would correspond to low alkalinities, the benefits of pH
buffering from the alkalinity will likely outweigh any benefits that might be achieved by
lowering the DIC of the finished water.
The impacts of water chemistry on the dissolution rate of PbO2 can also be used to
promote conditions with the slowest dissolution rates and consequently the best abilities to
maintain low lead concentrations in distribution systems. Not surprisingly, free chlorine is an
excellent inhibitor of the reductive dissolution of PbO2. Even very low concentrations of
residual free chlorine (0.2 mg/L as Cl2) can substantially inhibit dissolution. Orthophosphate can
also be effective at limiting the release of lead to water in systems that contain PbO2. The
benefits of orthophosphate addition may be limited to pH values near neutral. The inhibition of
PbO2 dissolution by orthophosphate does not necessarily require the precipitation of lead(II)
phosphate solids.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
APPENDIX A
EQUILIBRIUM CONSTANTS AND REACTIONS
#
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Table A.1
Equilibrium constants for aqueous species
Reaction
H2O = H+ + OHCO2(g) + H2O = H2CO3*
H2CO3* = 2H+ + CO32HCO3- = H+ + CO32H3PO4 = 3H+ + PO43H2PO4- = 2H+ + PO43HPO42- = H+ + PO43Pb2+ + H2O = PbOH+ + H+
Pb2+ + 2H2O = Pb(OH)20 + 2H+
Pb2+ + 3H2O = Pb(OH)3- + 3H+
Pb2+ + 4H2O = Pb(OH)42- + 4H+
Pb2+ + CO32- = PbCO30
Pb2+ + 2CO32- = Pb(CO3)22Pb2+ + CO3- + H+ = PbHCO3+
2Pb2+ + 3H2O = Pb2(OH)3+ + 3H+
3Pb2+ + 4H2O = Pb4(OH)42+ + 4H+
4Pb2+ + 4H2O = Pb4(OH)44+ + 4H+
HOCl + 2e- + H+ = Cl- + H2O
Pb4+ + 2e- = Pb2+
O2(aq) + 4H+ +4e- = 2H2O
2H+ +2e- = H2(aq)
22
Pb4+ + 3H2O = PbO32- + 6H+
23
Pb4+ + 4H2O = PbO44- + 8H+
a
values were calculated using the thermodynamic data in Table A.3.
Benjamin = (Benjamin 2002)
MINEQL = (Schecher and Mcavoy 1998)
Log K
-14.00
-1.46
-16.68
-10.33
-21.70
-19.56
-12.35
-7.71
-17.12
-28.06
-39.70
6.48
9.94
13.20
-6.40
-23.89
-19.99
50.20
28.64
86.00
Source
MINEQL
MINEQL
MINEQL
MINEQL
MINEQL
MINEQL
MINEQL
Benjamin
Benjamin
Benjamin
Benjamin
MINEQL
MINEQL
MINEQL
MINEQL
MINEQL
MINEQL
Benjamin
Benjamin
Benjamin
3.10
-23.04
-63.80
Benjamin
Calculateda
Calculated
73
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
74 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Table A.2.
Solubility products of select lead solids
Reaction
Log K
#
Solid
24
Massicot
PbO(s) + 2H+  Pb2+ + H2O
12.91
MINEQL
25
Litharge
PbO(s) + 2H+  Pb2+ + H2O
12.72
MINEQL
26
Pb(OH)2(s)
Pb(OH)2(s) + 2H+  Pb2+ + 2H2O
8.15
MINEQL
27
Cerussite
-13.13
Benjamin
-18.77
MINEQL
PbCO3(s)  Pb2+ + CO32+
Source
2+
Pb3(CO3)2(OH)2(s) + 2H  3Pb +
28
Hydrocerussite
29
Pb3(PO4)2(s)
Pb3(PO4)2(s)  3Pb2+ + 2PO43-
-44.50
Benjamin
30
PbHPO4(s)
PbHPO4(s) Pb2+ + PO43- + H+
-37.80
MINEQL
Hydroxyl-
Pb5(PO4)3OH(s) + H+  5Pb2+ +
-62.79
MINEQL
31
pyromorphite
2CO32- + 2H2O
3PO43- + H2O
32
Plattnerite
Pb(IV)O2(s) + 4H+  Pb4+ + 2H2O
-8.91
Calculateda
33
Scrutinyite
Pb(IV)O2(s) + 4H+  Pb4+ + 2H2O
-8.26
Calculated
a:
values were calculated using the thermodynamic data in Table A.3.
Benjamin = (Benjamin 2002)
MINEQL = (Schecher and Mcavoy 1998)
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Appendix A: Equilibrium Constants and Reactions | 75
Species
Table A.3
Chemical potentials for various aqueous species
G0f,i (J/mol)
Source
Pb3O4(s)
-601,200
Benjamin
Pb2O3(s)
-411,769
Pourbaix
α-PbO2(s)
-218,987
Pourbaix
β-PbO2(s)
-222,674
Risold
Pb2+
-24,309
Pourbaix
HPbO2-
-338,898
Pourbaix
Pb4+
302,498
Pourbaix
2-
-277,562
Pourbaix
PbO44-
-282,084
Pourbaix
H+
0.00
Benjamin
OH-
-157,300
Benjamin
H2O
-237,180
Benjamin
PbO3
Benjamin = (Benjamin 2002)
Pourbaix = (Pourbaix 1974)
Risold = (Risold et al. 1998)
The equilibrium constant for Reaction 22 was calculated by first determining the Gibbs
free energy of the reaction by the summation of the molar Gibbs free energies of formation ( G 0f ,i
) for each component ( i ), Equations A.1 and A.2. The Gibbs free energies of formation are
listed in Table A.3.
k
Gr   G 0f ,i N i
i 1
where Gr
G 0f ,i
Ni
(A.1)
= Gibbs free energy of reaction (r) , joules (J/mol)
= Gibbs free energy of formation of species (i), joules per mole (J/mol)
= stoichiometric coefficient
G r , 20   PbO 2   6   H    Pb 4   3   H 2O   131,481 J
3
The equilibrium constant can then be calculated by Equation A.3:
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
(A.2)
76 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
LogK eq  
where R
T
Gr
2.303RT
(A.3)
= equilibrium gas constant, joules per mole per Kelvin (J/mol•K)
= temperature, Kelvin (K)
Log ( K eq )  
131,481
  23.04
2.303RT
(A.4)
The equilibrium constant for Reaction 23 was calculated by the summation of the
chemical potentials in an analogous manner to Reaction 22.
The Gibbs free energy of reaction for Reaction 32 and 33 were determined by the
summation of free energies of formation (Table A.3), Equation A.1. The equilibrium constant
was then determined by Equation A.3.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
APPENDIX B
SUPPORTING INFORMATION ON LEAD(IV) OXIDE FORMATION
28 day
14 day
Relative Intensity
7 day
4 day
1 day
Cerussite PDF#01‐085‐1088
Hydrocerussite PDF#01‐073‐4362
Massicot PDF#00‐005‐0570
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure B.1 X-ray diffraction patterns of solids following reaction of massicot at pH 10 in
the absence of DIC with 20 mg Cl2/L free chlorine. (Reference patterns are listed at the
bottom; patterns for cerussite and hydrocerussite are included to show that they did not
form as intermediate solids).
77
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
78 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
pH 10
Relative Intensity
pH 8.5
pH 7.5
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure B.2 X-ray diffraction patterns of solids following 28 days of reaction of massicot
with 20 mg C/L DIC and 4 mg/L as Cl2 free chlorine together with reference patterns from
the ICDD database.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Appendix B: Supporting Information on Lead(IV) Oxide Formation | 79
28 day
Relative Intensity
7 day
1 day
Hydrocerussite PDF#01‐073‐4362
Massicot PDF#00‐005‐0570
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure B.3 X-ray diffraction patterns of solids following reaction of massicot at pH 10 in
the presence of 20 mg C/L DIC with 20 mg Cl2/L free chlorine together with reference
patterns from the ICDD database.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
80 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
28 day
7 day
Relative Intensity
1 day
Cerussite PDF#01‐085‐1088
Hydrocerussite PDF#01‐073‐4362
Massicot PDF#00‐005‐0570
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure B.4 X-ray diffraction patterns of solids following reaction of massicot at pH 7.5 in
the presence of 20 mg C/L DIC with 20 mg Cl2/L free chlorine together with reference
patterns from the ICDD database.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Appendix B: Supporting Information on Lead(IV) Oxide Formation | 81
28 day
relative intensity
7 day
1 day
Cerussite PDF#01‐085‐1088
Hydrocerussite PDF#01‐073‐4362
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Figure B.5 X-ray diffraction patterns of solids following reaction of hydrocerussite at pH
7.5 in the absence DIC with 20 mg Cl2/L free chlorine together with reference patterns
from the ICDD database.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
82 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
28 day
relative intensity
7 day
1 day
Cerussite PDF#01‐085‐1088
Hydrocerussite PDF#01‐073‐4362
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure B.6 X-ray diffraction patterns of solids following reaction of hydrocerussite at pH
10 in the absence DIC with 20 mg Cl2/L free chlorine together with reference patterns from
the ICDD database.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Appendix B: Supporting Information on Lead(IV) Oxide Formation | 83
28 day
relative intensity
7 day
1 day
Cerussite PDF#01‐085‐1088
Hydrocerussite PDF#01‐073‐4362
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Figure B.7 X-ray diffraction patterns of solids following reaction of cerussite at pH 7.5 in
the absence DIC with 20 mg Cl2/L free chlorine together with reference patterns from the
ICDD database.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
84 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
28 day
relative intensity
7 day
1 day
Cerussite PDF#01‐085‐1088
Hydrocerussite PDF#01‐073‐4362
Plattnerite PDF#01‐071‐4820
Scrutinyite PDF#04‐008‐7674
20
25
30
35
40
45
50
55
60
2θ (o)
Reprinted with permission from Wang et al. Copyright 2010 American Chemical Society.
Figure B.8 X-ray diffraction patterns of solids following reaction of cerussite at pH 10 in
the absence DIC with 20 mg Cl2/L free chlorine together with reference patterns from the
ICDD database.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
APPENDIX C
SUPPORTING INFORMATION ON LEAD(IV) OXIDE DISSOLUTION
RATES
EFFECT OF ORGANIC BUFFER ON PLATTNERITE DISSOLUTION
3-(N-morpholino)propanesulfonic acid (MOPS) was selected as the pH buffer in the
experiments without dissolved inorganic carbon (DIC) at pH 7.6. A previous study suggested
that organic buffers, like 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES), and 2(N-morpholino)ethanesulfonic acid (MES), could significantly enhance the dissolution of PbO2,
probably by acting as reductants for PbO2 (Xie et al. 2010). Therefore the effect of MOPS on the
dissolution of plattnerite was evaluated using both batch reactors and flow-through reactors.
Batch experiments were conducted at room temperature (21 ± 1 ºC) in 500-mL
polypropylene batch reactors. In each experiment aliquots of a 0.5 M NaHCO3 stock solution
were used to adjust the DIC concentration to 50 mg C/L. Then the desired MOPS concentration
(0, 1, or 10 mM) was provided by addition of an aliquot of 0.5 M MOPS stock solution. The pH
was then adjusted to 7.6 by addition of concentrated HNO3 or freshly prepared 0.5 M NaOH
solution. Finally PbO2 solids were added to provide a solid loading of 50 mg/L. Experiments
were conducted up to 24 hours. Samples were collected at different time intervals, filtered
through 0.22 μm polyether sulfone (PES) syringe filters, acidified to 2% HNO3, and preserved
for dissolved lead analysis. Each experimental condition was run in duplicate.
Flow-through experiments were performed using the continuously stirred tank reactors
(CSTRs). Experiments were conducted at pH 7.6 with 50 mg C/L DIC and 10 mM NaNO3 in the
absence and presence of 1 mM MOPS for 24 hours. Each experimental condition was run in
duplicate.
In the batch experiments the presence of MOPS enhanced the dissolution of plattnerite,
and the dissolved lead concentrations after 24 hours of reaction increased with increasing MOPS
concentrations (Figure C.1). While around 7 ug/L dissolved lead was released in the absence of
MOPS, the dissolved lead concentration increased to about 20 ug/L with 1 mM MOPS and to
200 ug/L with 10 mM MOPS. In the flow-through experiments, the steady-state effluent lead
concentration without MOPS was about 30 nM, and it increased to 50 nM with 1 mM MOPS
present (Figure C.2). Therefore the dissolution rate of plattnerite in the presence of 1 mM MOPS
was only 1.7 times higher than without MOPS. Although the presence of 1 mM MOPS slightly
promoted the dissolution of plattnerite, this slight enhancement was acceptable for the
experiments on the effects of DIC because the effect of DIC was much greater than that of
MOPS. The MOPS was critical to maintaining a stable pH while probing the effects of DIC.
85
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
86 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
1000 0 mM
1 mM
Dissolved Pb (µg/L)
10 mM
100 10 1 0
5
10
15
20
25
time (h)
Figure C.1 Dissolved lead concentration from batch reactors over time at pH 7.6 with 50
mg C/L DIC and different concentrations of MOPS.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
Appendix C: Supporting Information on Lead(IV) Oxide Dissolution Rate | 87
100
90
Dissolved Pb (nM)
80
70
60
50
40
30
20
10
0
0
10
20
30
40
50
Residence Times (t/τ)
Figure C.2 Effluent lead concentrations from CSTRs as a function of the number of
hydraulic residence times (τ = 30 min) at pH 7.6 with 50 mg C/L DIC in the absence (□
and △) and presence of 1 mM MOPS (■ and ▲). Duplicate experiments (represented by
squares and triangles) were conducted for each condition.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
88 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
EFFLUENT LEAD CONCENTRATIONS OVER TIME
10 300
no I‐
(a)
Dissolved Pb (nM)
250
9
200
250
9
200
150
5
0
1600
10
(d)
20
30
40
6
5
0
0
10 1600
1400
10
(e)
20
30
40
10 μM I‐
1200
6
50
5
0
50
0
10 1600
1400
9
7
100
50
50
5 μM I‐
8
7
100
0
9
150
7
6
250
8
150
100
10
2 μM I‐
(c)
200
8
50
10
(f)
20
30
40
50
20 μM I‐
10
1400
9
1200
9
1200
1000
8 1000
8 1000
800
800
800
pH
Dissolved Pb (nM)
10 300
1 μM I‐
(b)
pH
300
600
7
7
600
400
6
200
7
600
400
6
200
5
0
0
10 20 30 40 50
Residence Times (t/tres)
8
400
6
200
5
0
0
10 20 30 40 50
Residence Times (t/tres)
5
0
0
10 20 30 40 50
Residence Times (t/tres)
Figure C.3 Effluent lead concentrations (■ and ▲) and pH (□ and △) from CSTRs as a
function of the number of hydraulic residence times (τ = 30 min) at pH 7.6 with different
iodide concentrations. Panels (a) – (f) represents the iodide concentration of 0, 1, 2, 5, 10, 20
μM respectively. Duplicate experiments (represented by squares and triangles) were
conducted for each condition.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
REFERENCES
American Society of Civil Engineers (2005) Report Card for America's Infrastructure.
AWWA (2005) Managing change and unintended consequences: Lead and Copper Rule
corrosion control treatment, American Water Works Association, Denver, CO.
Awwa Research Foundation (2007) Distribution System Water Quality, Strategic Initiative Plan,
Awwa Research Foundation, Denver, CO.
Benjamin, M.M. (2002) Water chemistry, McGraw-Hill, Boston.
Boyd, G., Dewis, K., Sandvig, A.M., Hendrickson, S., Vetrovs, A., Korshin, G.V., McFadden,
M., Reiber, S.H., Giani, R. and Otto, S. (2007) Effect of changing disinfectants on
distribution system lead and copper release, Proc. AWWA Water Quality Technology
Conference, Charlotte, NC.
Cantor, A.F. (2006) Diagnosing corrosion problems through differentiation of metal fractions.
Journal American Water Works Association 98(1), 117-126.
Cantor, A.F., Park, J.K. and Vaiyavatjamai, P. (2003) Effect of chlorine on corrosion in drinking
water systems. Journal American Water Works Association 95(5), 112-123.
Cerrato, J.M., Reyes, L.P., Alvarado, C.N. and Dietrich, A.M. (2006) Effect of PVC and iron
materials on Mn(II) deposition in drinking water distribution systems. Water Research
40(14), 2720-2726.
Churchill, D.M., Mavinic, D.S., Neden, D.G. and MacQuarrie, D.M. (2000) The effect of zinc
orthophosphate and pH-alkalinity adjustment on metal levels leached into drinking water.
Canadian Journal of Civil Engineering 27(1), 33-43.
Clesceri, L.S., Greenberg, A.E. and Eaton, A.D. (eds) (1999) Standard Methods for the
Examination of Water and Wastewater, American Public Health Association, American
Water Works Association, Water Environment Federation, Washington, DC.
Dodrill, D.M. and Edwards, M. (1995) Corrosion control on the basis of utility experience.
Journal American Water Works Association 87(7), 74-85.
Dryer, D.J. and Korshin, G.V. (2007) Investigation of the reduction of lead dioxide by natural
organic matter. Environmental Science & Technology 41(15), 5510-5514.
Edwards, M. (2004) Controlling corrosion in drinking water distribution systems: a grand
challenge for the 21st century. Water Science and Technology 49(2), 1-8.
Edwards, M. and Dudi, A. (2004) Role of chlorine and chloramine in corrosion of lead-bearing
plumbing materials. Journal American Water Works Association 96(10), 69-81.
Edwards, M., Jacobs, S. and Dodrill, D. (1999) Desktop guidance for mitigating Pb and Cu
corrosion by-products. Journal American Water Works Association 91(5), 66-77.
Edwards, M. and McNeill, L.S. (2002) Effect of phosphate inhibitors on lead release from pipes.
Journal American Water Works Association 94(1), 79-90.
Fuge, R. and Johnson, C. (1986) The geochemistry of iodine-a review. Environmental
Geochemistry and Health 8(2), 31-54.
Giammar, D., Nelson-Nguyen, K., Xie, Y. and Noel, J. (2008) Role of phosphate in mitigating
lead release from corrosion products, Water Quality Technology Conference, Cincinnati,
OH.
Giammar, D.E., Nelson, K.S., Noel, J.D. and Xie, Y. (2010) Water chemistry effects on
dissolution rates of lead corrosion products, Water Research Foundation, Denver, CO.
89
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
90 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Glasstone, S. (1922) Physical chemistry of the oxide of lead. Part V. The electromotive
behaviour of lead dioxide. Journal of the Chemical Society 122, 1469-1480.
Hering, J.G. and Stumm, W. (1990) Oxidative and reductive dissolution of minerals. Reviews in
Mineralogy and Geochemistry 23(1), 427-465.
Hozalski, R.M., Esbri-Amador, E. and Chen, C.F. (2005) Comparison of stannous chloride and
phosphate for lead corrosion control. Joural of the American Water Works Association
97(3), 89-103.
Lin, Y.-P. and Valentine, R.L. (2008a) The release of lead from the reduction of lead oxide
(PbO2) by natural organic matter. Environmental Science & Technology 42(3), 760-765.
Lin, Y.-P., Washburn, M. and Valentine, R.L. (2008) Reduction of lead oxide (PbO2) by iodide
and formation of iodoform in PbO2/I-/NOM system. Environmental Science &
Technology 42(8), 2919-2924.
Lin, Y.P. and Valentine, R.L. (2008b) The release of lead from the reduction of lead oxide
(PbO2) by natural organic matter. Environmental Science & Technology 42(3), 760-765.
Lin, Y.P. and Valentine, R.L. (2009) Reduction of Lead Oxide (PbO2) and Release of Pb(II) in
Mixtures of Natural Organic Matter, Free Chlorine and Monochloramine. Environmental
Science & Technology 43(10), 3872-3877.
Lin, Y.P. and Valentine, R.L. (2010) Reductive Dissolution of Lead Dioxide (PbO2) in Acidic
Bromide Solution. Environmental Science & Technology 44(10), 3895-3900.
Liu, H., Korshin, G.V., Ferguson, J.F. and Jiang, W. (2006) Key parameters and kinetics of
oxidation of lead(II) solid phases by chlorine in drinking water. Water Practice and
Technology 1(4).
Liu, H.Z., Korshin, G.V. and Ferguson, J.F. (2008) Investigation of the kinetics and mechanisms
of the oxidation of cerussite and hydrocerussite by chlorine. Environmental Science &
Technology 42(9), 3241-3247.
Lytle, D.A. and Schock, M.R. (2000) Impact of stagnation time on metal dissolution from
plumbing materials in drinking water. Journal of Water Supply Research and
Technology-Aqua 49(5), 243-257.
Lytle, D.A. and Schock, M.R. (2005) Formation of Pb(IV) oxides in chlorinated water. Journal
American Water Works Association 97(11), 102-114.
Lytle, D.A., Schock, M.R. and Scheckel, K. (2009) The Inhibition of Pb(IV) Oxide Formation in
Chlorinated Water by Orthophosphate. Environmental Science & Technology 43(17),
6624-6631.
McNeill, L.S. and Edwards, M. (2002) Phosphate inhibitor use at US utilities. Journal American
Water Works Association 94(7), 57-63.
National Academy of Sciences (1980) Drinking water and health, National Academy Press,
Washington D.C.
Noel, J.D. and Giammar, D.E. (2007) The influence of water chemistry on dissolution rates of
lead corrosion products, 2007 Water Quality Technology Conference, Charlotte, NC.
Nour, S., Prevost, M., Cartier, C. and Laroche, L. (2007) Impact of sampling flow rate, flushing,
and faucet aerator on dissolved and particualte lead concentrations at consumer's tap,
Proc. AWWA Water Quality Technology Conference, Charlotte, NC.
Pokrovsky, O.S. and Schott, J. (2002) Surface chemistry and dissolution kinetics of divalent
metal carbonates. Environmental Science & Technology 36(3), 426-432.
Pourbaix, M. (1974) Atlas of Electrochemical Equilibrium in Aqueous Solutions, National
Association of Corrosion Engineers, Houston, TX.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
References | 91
Rajasekharan, V.V., Clark, B.N., Boonsalee, S. and Switzer, J.A. (2007) Electrochemistry of free
chlorine and monochloramine and its relevance to the presence of Pb in drinking water.
Environmental Science & Technology 41, 4252-4257.
Renner, R. (2004) Plumbing the depths of D.C.'s drinking water crisis. Environmental Science &
Technology 38(12), 224A.
Renner, R. (2006) Chloramine's effect on lead in drinking water. Environmental Science &
Technology 40(10), 3129-3130.
Renner, R. (2007) Lead levels high in Canadian tap water. Environmental Science &
Technology.
Risold, D., Nagata, J.I. and Suzuki, R.O. (1998) Thermodynamic description of the Pb-O system.
Journal of Phase Equilibria 19(3), 213-233.
Samson, S.D., Stillings, L.L. and Eggleston, C.M. (2000) The depletion and regeneration of
dissolution-active sites at the mineral-water interface: I. Fe, Al, and In sesquioxides.
Geochimica Et Cosmochimica Acta 64(20), 3471-3484.
Schecher, W.D. and Mcavoy, D.C. (1998) MINEQL: A chemical equilibrium modeling system,
version 4.5. Howell, ME, Environmental Research Software.
Schock, M.R. (1989) Understanding Corrosion Control Strategies for Lead. Journal American
Water Works Association 81(7), 88-100.
Schock, M.R., S.M. Harmon, J. Swertfeger, and R. Lohmann (2001) Tetravalent lead: A Hitherto
unrecognized control of tap water lead contamination. Proc. AWWA Water Quality
Technology Conference, Nashville, Tennessee, November 11–15.
Schock, M. and Giani, R. (2004) Oxidant/Disinfectant Chemistry and Impacts on Lead
Corrosion, Proc. AWWA Water Quality Technology Conference, San Antonio, TX.
Schock, M.R., Lytle, D.A., Sandvig, A.M., Clement, J. and Harmon, S.M. (2005a) Replacing
polyphosphate with silicate to solve lead, copper, and source water iron problems.
Journal American Water Works Association 97(11), 84-93.
Schock, M.R., Scheckel, K., DeSantis, M. and Gerke (2005b) Mode of Occurrence, Treatment,
and Monitoring Significance of Tetravalent Lead, Proc. AWWA Water Quality
Technology Conference, Quebec City.
Schock, M.R., Hyland, R.N. and Welch, M.M. (2008) Occurrence of contaminant accumulation
in lead pipe scales from domestic drinking-water distribution systems. Environmental
Science & Technology 42(12), 4285-4291.
Schock, M.R. and Lytle, D.A. (2010) Internal Corrosion and Deposition Control. In Water
Quality and Treatment. Edzwald, J.K. (ed), McGraw-Hill, Inc., New York, New York.
Shi, Z. and Stone, A.T. (2009a) PbO2(s, Plattnerite) Reductive Dissolution by Aqueous
Manganous and Ferrous Ions. Environmental Science & Technology 43(10), 3596-3603.
Shi, Z. and Stone, A.T. (2009b) PbO2(s, Plattnerite) Reductive Dissolution by Natural Organic
Matter: Reductant and Inhibitory Subfractions. Environmental Science & Technology
43(10), 3604-3611.
Stone, A.T. and Morgan, J.J. (1984) Reduction and dissolution of manganese(III) and
manganese(IV) oxides by organics. 1. Reaction with hydroquinone. Environmental
Science & Technology 18(6), 450-456.
Stumm, W. and Morgan, J.J. (1996) Aquatic chemistry, John Wiley & Sons, Inc., New York,
NY.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
92 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
Switzer, J.A., Rajasekharan, V.V., Boonsalee, S., Kulp, E.A. and Bohannan, E.W. (2006)
Evidence that monochloramine disinfectant could lead to elevated Pb levels in drinking
water. Environmental Science & Technology 40(10), 3384-3387.
U.S. Environmental Protection Agency (2007a) Elevated Lead in D.C. Drinking Water – A
Study of Potential Causative Events, Final Summary Report, U.S. Environmental
Protection Agency.
U.S. Environmental Protection Agency (2007b) Summary of Research Relevant to the D.C. Lead
Issue, Washington, D.C.
Vasquez, F.A., Heaviside, R., Tang, Z.J. and Taylor, J.S. (2006) Effect of free chlorine and
chloramines on lead release in a distribution system. Journal American Water Works
Association 98(2), 144-154.
Volk, C., Wood, L., Johnson, B., Robinson, J., Zhu, H. and Kaplan, L. (2002) Monitoring
dissolved organic carbon in surface and drinking waters. Journal of Environmental
Monitoring 4, 43-47.
Wang, Y., Xie, Y., Li, W. and Giammar, D.E. (2010) Formation of lead(IV) oxide corrosion
products from lead(II) compounds. Environmental Science & Technology 44(23), 89508956.
Xie, L. and Giammar, D.E. (2007) Equilibrium solubility and dissolution rate of the lead
phosphate chloropyromorphite. Environmental Science & Technology 41(23), 80508055.
Xie, Y. and Giammar, D.E. (2011) Effects of flow and water chemistry on lead release rates
from pipe scales. Water Research 45(19), 6525-6534.
Xie, Y.J., Wang, Y., Singhal, V. and Giammar, D.E. (2010) Effects of pH and Carbonate
Concentration on Dissolution Rates of the Lead Corrosion Product PbO2. Environmental
Science & Technology 44(3), 1093-1099.
Zhang, Y., Griffin, A., Rahman, M., Camper, A., Baribeau, H. and Edwards, M. (2009) Lead
Contamination of Potable Water Due to Nitrification. Environmental Science &
Technology 43(6), 1890-1895.
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
ABBREVIATIONS
A
AWWA
specific surface area
American Water Works Association
BDL
BET
below detection limit
Brunauer Emmett and Teller adsorption isotherm
°C
C(t)
Ceff
Ceq
Cinf
cm2
Css
CSTR
CT
degrees Celsius
concentration as a function of time
effluent concentration
predicted equilibrium concentration
influent concentration
square centimeters
steady-state concentration
Continuously-stirred tank reactor
total dissolved inorganic carbon concentration
DIC
dissolved inorganic carbon
Ea
EDX
activation energy
energy dispersive X-ray analysis
f(ΔG)
ft/s
FTIR
generic expression of function of Gibbs free energy effect on rate
feet per second
Fourier transform infrared spectroscopy
g/L
Gf°i
ΔG
grams per liter
Gibbs free energy of formation of species i
Gibbs free energy of reaction
IAP
ICDD
ICP-MS
ion activity product
International Centre for Diffraction Data
inductively coupled plasma mass spectrometry
k
k0
Ksp
dissolution rate constant
intrinsic dissolution rate constant
solubility product
L
LCR
liters
Lead and Copper Rule
M
m/s
m2/g
moles per liter
meters per second
square meters per gram
93
©2012 Water Research Foundation. ALL RIGHTS RESERVED.
94 | Lead(IV) Oxide Formation and Stability in Drinking Water Distribution Systems
mg C/L
mg Cl2/L
mg P/L
mg/L
mL
mM
mol/min·m2
MOPS
milligrams of carbon per liter
milligrams of free chlorine per liter
milligrams of phosphorus per liter
milligrams per liter
milliliters
millimoles per liter
moles per minute per square meter
3-(N-morpholino) propanesulfonic acid
ng/L
nH+
ni
Ni
nm
nM
NOM
nanograms per liter
order of reaction with respect to H+
order of reaction with respect to species i
stoichiometric coefficient of species i
nanometers
nanomoles per liter
natural organic matter
ORP
oxidation-reduction potential
PAC
PCO2
PES
PM
ppb
ppt
project advisory committee
partial pressure of carbon dioxide
polyethersulfone
Project Manager
parts per billion
parts per trillion
R
rpm
ideal gas constant
revolutions per minute
s.t.
SEM
SI
SIMS
stagnation time
scanning electron microscopy
saturation index
secondary ion mass spectrometry
T
t
tres
temperature
time
hydraulic residence time
U.S.EPA
United States Environmental Protection Agency
XRD
X-ray diffraction
μ
μg/L
μm
Ω
chemical potential
micrograms per liter
micrometers
empirical coefficient to account for non-elementary reactions
©2012 Water Research Foundation. ALL RIGHTS RESERVED.